91
i Natural gas conversion The Reforming and Fischer-Tropsch processes by Sarah Lögdberg Hugo A. Jakobsen TKP 4145 Reactor technology Compendium

Natural Gas Conversion

  • Upload
    asiteng

  • View
    89

  • Download
    4

Embed Size (px)

DESCRIPTION

-

Citation preview

Page 1: Natural Gas Conversion

i

Natural gas conversion

The Reforming and Fischer-Tropsch processes

by

Sarah Lögdberg

Hugo A. Jakobsen

TKP 4145 Reactor technology

Compendium

Page 2: Natural Gas Conversion

ii

Prelude

This compendium consists of two parts. The first part deals with natural gas reforming in

general, whereas the second part treats the modelling of trickle bed reactors and the Fischer-

Tropsch synthesis.

Page 3: Natural Gas Conversion

iii

Table of contents

Prelude .................................................................................................. ii

Reforming of natural gas .................................................................... 1

1. Introduction ........................................................................................................................ 2

2. Different natural gas reforming concepts ........................................................................... 3

2.1 Reactions and thermodynamics: ................................................................................... 5

2.2 Steam reforming (tubular reforming) ........................................................................... 8

2.3 Adiabatic pre-reforming ............................................................................................. 13

2.4 Partial oxidation ......................................................................................................... 14

2.5 Autothermal reforming and secondary reforming ...................................................... 15

2.6 Heat exchange reforming ........................................................................................... 18

3. Different reforming concepts for different applications................................................... 21

3.1 Introduction ................................................................................................................ 21

3.2 Methanol synthesis ..................................................................................................... 23

3.3 Ammonia synthesis .................................................................................................... 24

3.4 Fischer-Tropsch synthesis .......................................................................................... 24

4. New reforming concepts under development ................................................................... 26

5. Some notes on modeling of CPOX and ATR .................................................................. 28

Modeling of trickle-bed reactors ...................................................... 31

1. Fixed-bed reactors vs trickle-bed reactors ....................................................................... 32

2. Dispersion model equations for multiphase systems ....................................................... 34

3. Special issues in modeling of trickle-bed reactors ........................................................... 38

3.1 Introduction ................................................................................................................ 38

3.2 Catalyst wetting and relating kapp to k ........................................................................ 39

3.3 Liquid hold-up and mass transfer ............................................................................... 41

3.4 Heat transfer ............................................................................................................... 44

3.5 Effectiveness factor .................................................................................................... 45

3.6 Backmixing ................................................................................................................ 46

3.6.1 One-parameter models ............................................................................................ 46

3.6.2 Two-parameter model ............................................................................................. 49

Page 4: Natural Gas Conversion

iv

3.7 Conclusion .................................................................................................................. 51

The Fischer-Tropsch process ............................................................................................... 52

4.1 Introduction ................................................................................................................ 52

4.2 Different reactor types and reaction conditions ......................................................... 54

4.2.1 HTFT ....................................................................................................................... 55

4.2.2 LTFT ....................................................................................................................... 56

4.3 Scope of the work ....................................................................................................... 58

4.4 Modeling of the trickle-bed FT-reactor ...................................................................... 59

4.4.1 Wang et al. [2003] ................................................................................................... 61

4.4.2 Jess et al. [1999] ...................................................................................................... 70

4.5 Modeling of the slurry phase reactor .......................................................................... 73

4.5.1 Different flow regimes and estimation of gas hold-up ............................................ 73

4.5.2 Mass transfer ........................................................................................................... 76

4.5.3. Mixing .................................................................................................................... 77

4.5.4 Heat transfer ............................................................................................................ 78

4.5.5 The model ................................................................................................................ 78

4.5.6 The results ............................................................................................................... 81

5. Conclusion/discussion ...................................................................................................... 83

References ........................................................................................... 85

Page 5: Natural Gas Conversion

1

Reforming of natural gas - The different concepts

Page 6: Natural Gas Conversion

2

1. Introduction

The largest natural gas reserves are located in the former USSR, Iran and Qatar. There are

also large reserves in the North Sea. About 40 % of the resources are in the former USSR.

[www.gasforeningen.se] In 1993, Norway produced 1.1 % of the world natural gas production

[Kirk-Othmer, 1995]. Natural gas stands for approximately 21 % of the world’s energy

consumption [www.iea.org].

Except for being used as fuel, natural gas is the most common feedstock for hydrogen

production or syngas production for synthesis of base chemicals (such as methanol and

ammonia), oil refining, and in many other industrial applications (iron ore reduction,

hydrogenation of fats, production of oxo alcohols). Of the total hydrogen production by

catalytic reforming, 50 % is used for ammonia production, 10 % for methanol production and

35 % for hydrotreatment in petrochemical processes (hydrocracking, hydrodesulfurization

etc.) [Ullman’s, 1985].

Syngas may however be produced from a range of feedstocks (naphtha, coal, biomass etc.).

For heavy hydrocarbons as feedstock, partial oxidation with steam and oxygen is used for

syngas production. Also for solid feedstocks, such as coal and biomass, partial oxidation with

steam and oxygen is used for the syngas production, but in this case it is often referred to as

gasification. When producing syngas from natural gas or light hydrocarbons, steam reforming

is often the preferred route if a high H2/CO ratio is desired. Steam reforming can be

performed in the presence of oxygen or CO2. A combination of steam reforming and partial

oxidation is called autothermal reforming, in which the endothermic and exothermic reactions

are coupled [Moulijn et al., 2003].

For light hydrocarbons, partial oxidation is usually not economical since it requires an

expensive cryogenic air separation unit. However, for special applications, a high CO/H2 ratio

in the syngas is desired, which can be accomplished by partial oxidation. Partial oxidation

may be carried out with or without a catalyst [Moulijn et al., 2003].

It is desirable to operate at higher pressures, if the syngas is to be used for production of

chemicals such as methanol, ammonia or Fischer-Tropsch fuels. This is because these fuel

synthesis operate at elevated pressures, and hence a pressurized syngas removes the cost of

compressing. Furthermore, plants operating at elevated pressures can have higher capacities

due to their smaller gas volumes, which in turn means lower investment costs [Ullman’s,

1985].

The gasification process may be conducted with or without a catalyst. The use of a catalyst

makes it possible to come closer to the equilibrium conversion, but catalysts can only be used

for gaseous feeds or for distillate liquid feeds.

In this report, the state-of-the-art of reforming of natural gas into synthesis gas will be

described. Different commercial concepts will be discussed and compared, as well as

concepts being developed at present, not yet commercialised.

Page 7: Natural Gas Conversion

3

2. Different natural gas reforming concepts

Table 1 shows the approximate molar H2/CO ratios that are achieved upon different natural

gas reforming concepts. Wilhelm et al. [2001] summarized the advantages and disadvantages

with the major natural gas syngas production concepts in Table 2.

Table 1. Ways to increase/reduce H2/CO ratios, and approximate H2/CO ratios from different concepts [Wilhelm

et al., 2001].

Page 8: Natural Gas Conversion

4

Table 2. Advantages and disadvantages of several natural gas reforming concepts. SMR = steam methane

reforming, ATR = autothernal reforming, POX = partial oxidation. aSMR followed by oxygen-blown secondary

reforming [Wilhelm et al., 2001].

Page 9: Natural Gas Conversion

5

2.1 Reactions and thermodynamics:

The standard enthalpies of reaction (at 298 K) are given in brackets. The most important

reactions in steam reforming (SR) of methane are [Moulijn et al., 2003]:

1. CH4 + H2O CO + 3H2 (+206 kJ/mol)

2. CO + H2O CO2 + H2 (water-gas-shift, WGS, -41 kJ/mol)

3. CH4 + CO2 2CO + 2H2 (dry reforming, +247 kJ/mol)

The last one is called “dry reforming” because it can be used to produce a CO-rich syngas

without steam as reactant.

These main reactions may be accompanied by coke formation according to:

4. CH4 C + 2H2 (decomposition of methane, +75 kJ/mol)

5. 2CO C + CO2 (the Boudouard reaction, -173 kJ/mol)

Reactions (4) and (5) are reversible. At temperatures above 650 ºC, also higher hydrocarbons

may be thermally cracked into coke, and this reaction is irreversible. [Dybkjaer, 1995] The

carbon formation will be discussed more deeply in 2.2 Steam reforming.

In the presence of oxygen, the following reactions may occur [Moulijn et al., 2003]:

6. CH4 + ½O2 CO + 2H2 (-36 kJ/mol)

7. CH4 + 2O2 CO2 + 2H2O (-803 kJ/mol)

8. CO + ½O2 CO2 (-284 kJ/mol)

9. H2 + ½O2 H2O (-242 kJ/mol)

Since the reaction with methane and steam (1) is highly endothermic, and the reactions

involving O2 (6 – 9) are highly exothermic, the production of syngas from methane can me

allothermal or autothermal. Allothermal means that the required heat for the reaction is

produced outside of the reactor. This is the case for pure steam reforming, or when only little

amount of O2 is present. Autothermal means that all heat required for the steam reforming

reaction is formed inside of the reactor by means of exothermic reactions with O2, and hence

no external heating is needed. It is the steam/oxygen ratio that determines the mode of

operation [Moulijn et al., 2003]. When the reaction is exothermal it takes place in adiabatic

reactors.

Dybkjaer [1995] remarked that also higher hycrocarbons are present in natural gas, and that

they will react according to the following endothermal reaction:

CnHm + nH2O nCO + (n+m/2)H2 (10)

This reaction is irreversible and proceeds to full conversion. The overall heat of reaction of

(1), (2) and (10) may be positive, zero or negative, depending on the process conditions. At

low steam/carbon (S/C) ratios, and at low catalyst exit temperatures, the overall reaction is

only slightly endothermal or even exothermal, if the feed contains high concentrations of

Page 10: Natural Gas Conversion

6

higher hydrocarbons. This is due to that the CO formed in reaction (10) is converted into CH4

according to the reverse of reaction (1). In this case, the process may be carried out without

external heating, e.g. in an adiabatic pre-reformer. However, if a syngas with a low methane

content is desired, a high exit temperature is required and the overall heat of reaction will be

endothermal, and external heat is necessary [Dybkjaer, 1995].

As shown in Figure 1, at high temperatures (> 800 ºC) the equilibrium H2/CO-ratio of the

syngas reaches a value of 3/1 in case of pure steam reforming (1), and 2/1 in case of partial

oxidation (POX) (6), respectively. These values are valid if the feed has the stoichiometric

CH4/H2O and CH4/O2 ratios for the steam reforming and partial oxidation reactions,

respectively. The data in Figure 1 are obtained if respect is taken to all reactions 1 – 3 for the

left picture (a), and 1 – 3 and 6 – 9 for right picture (b). Their similar appearance is due to the

fact that the same reactions actually occur in both cases, except for the reactions in which O2

are involved which only occur in the partial oxidation case [Moulijn et al., 2003].

Figure 1. Effect of temperature on equilibrium composition at 1 bar in steam reforming of methane (a), and

partial oxidation of methane (b). H2O/CH4 = 1/1 in (a), O2/CH4 = 0.5/1 in (b) [Moulijn et al., 2003].

The peak in CO2 concentration is explained by that formation of CO2 is favoured at low

temperatures, since CO2 is only formed in exothermal reactions. At higher temperatures, the

CO2 is consumed in the endothermal reforming reactions, since these reactions are favoured

by higher temperatures [Moulijn et al., 2003].

Figure 2 shows the pressure dependence of the equilibrium composition in case of pure steam

reforming with stoichiometric feed. A higher pressure shifts the steam reforming equilibrium

towards the reactants since they compose fewer molecules. At 30 bar, a temperature of 1400

K is needed to reach an equilibrium in which only the products CO and H2 exist. However,

the maximum temperatures used in industry for steam reforming are around 1200 K due to

reactor material constraints [Moulijn et al., 2003].

Page 11: Natural Gas Conversion

7

Figure 2. Effect of temperature and pressure on equilibrium composition in steam reforming of methane.

H2O/CH4 = 1/1. (a) H2 and CH4, (b) CO, H2O and CO2 [Moulijn et al., 2003].

Since most syngas applications, such as methanol and ammonia synthesis, require high

pressures, the syngas production is often performed at high pressure in order to avoid an

expensive compression and to make the reformer size small. A high pressure is not favourable

for equilibrium of the syngas production, and hence a high temperature and steam in excess is

used as compensation, otherwise the methane conversion is too low. Figures 3 shows the

effect of pressure and the effect of steam/methane ratio on unconverted methane (the methane

slip), respectively. The most economical operation is as high pressure and temperature as

possible, the tube material setting the limits. [Moulijn et al., 2003] In other cases, the pressure

is set by the requirements of downstream separation or purification processes, such as PSA

units, membranes or cryogenic units. [Dybkjaer, 1995].

Figure 3. Effect of pressure (left) and steam/methane ratio (right) on unconverted methane as a function of

temperature [Moulijn et al., 2003]

Page 12: Natural Gas Conversion

8

2.2 Steam reforming (tubular reforming)

The overall steam reforming (SR) is highly endothermic and it is carried out at high

temperature (900 ºC) and at pressures between 15 and 30 bar [Moulijn et al., 2003] over a

Ni/Al2O3 catalyst [Bharadwaj and Schmidt, 1995]. The composition of the gas at the reactor

outlet reflects the equilibria of reaction (1) and (2) above. In some cases, it is advantageous to

add CO2 at the inlet of the reformer. This is done in order to save hydrocarbon feedstock and

decrease the H2/CO ratio in the product gas. The CO2 coming out from the reformer is then

recycled, and in some cases CO2 is also imported. CO2 then is supposed to react to form CO

via the reverse WGS-reaction, and this is favored at low S/C ratios. However, low S/C ratios

leads to high methane concentrations in the outlet. To compensate for this, a higher

temperature can be used [Dybkjaer, 1995].

The Ni-catalyst is needed since methane is a very thermodynamically stable molecule even at

high temperatures. The steam reformer consists of two sections – a convection section and a

radiant section (see Figure 4). In the convection section, methane and steam are preheated to

500 ºC [Ullman’s, 1985] by heat exchange with the hot flue gases from the fuel combusted in

the reformer furnace, and in the radiant section the reforming reactions take place in the

catalyst filled tubes which are hanging in rows inside of the reformer furnace [Moulijn et al.,

2003]. The process gas is heated gradually to approximately 800 C inside of the tubes

[Ullman’s, 1985]. All tubular reformers use catalyst inside the tubes in order to reduce the

operating temperature. This is important in order to reduce the tube stresses resulting from

high pressure and high temperatures [Froment and Bischoff, 1990]. HP-steam is also

produced in the convection zone, which may be converted to electricity which is used for

compression of the produced syngas in case of methanol or ammonia synthesis. Steam for the

reforming is produced from the heat in the product gas out from the reformer. One furnace

can contain 500 – 600 tubes with inner diameters of 70 –130 mm and lengths of 7 – 12 m

[Moulijn et al., 2003]. The wall thickness of the tubes is between 10 – 20 mm [Ullman’s,

1985]. The tubes have small diameters in order to achieve the highest possible heat flux to the

catalyst, and hence to achieve the highest possible capacity for a given amount of catalyst

[Froment and Bischoff, 1990].

The heat needed in the tubular reformer is the enthalpy difference between inlet and exit gas.

The heat duty consists of the heat of the reaction and the heat needed to bring the products to

the exit temperature. Approximately 50 % of the heat produced by combustion in the burners

is transferred to the process gas. The other 50 % exits the system in the hot flue gases from

the burners, and this heat is recovered in the convection section as described above. The

overall thermal efficiency of the reformer may approach 95 % [Dybkjaer, 1995]. In the case

where there is no need for surplus energy, and hence production of HP-steam is not desirable,

one solution is to install smaller tubes inside the reformer tube. The catalyst is placed in the

space between the main and the smaller tubes. After being preheated, the steam and methane

enter the catalyst-filled space, where it is heated and reforming takes place as usual. At the

end of the reformer tube, the gas then enters the smaller tubes and transmits some heat to the

catalyst bed before being discharged at the top. This reduces the amount of heat transmitted

by the reformer tube and hence the number of tubes and their surface area by approximately

20 % [Ullman’s, 1985].

Page 13: Natural Gas Conversion

9

Figure 4. Schematic picture of a steam reformer, showing the radiation and convection sections [Moulijn et al.,

2003].

Methane reforming can be described by a first-order reaction, irrespective of pressure. At high

temperatures the overall rate can be limited by pore diffusion, but at low temperatures the

molecular diffusion rate is much higher than the reaction rate so that the catalyst activity can

be fully used [Ullman’s, 1985]. According to Dybkjaer [1995], the overall rate in steam

reforming is limited by the heat transfer, at high temperatures. The Ni-catalyst is often in the

form of thick-walled Raschig rings, with 16 mm in diameter and height, and a 6 – 8 mm hole

in the middle. If the heat load per unit area is too high, the limits of such catalysts will be

reached, and hence smaller particles will be necessary in order to make use of more of the

catalyst. Smaller particles will however lead to increased pressure drop. Therefore, special

packing shapes such as spoked wheels or rings with several holes have been developed for

these smaller particles [Ullman’s, 1985].

The Ni-catalyst is poisoned by sulfur, which is present in practically all gaseous feedstocks,

why desulfurization is a necessary step prior to reforming (see Figure 4). The desulfurization

is often made with zinc-oxide absorption beds working at 350 – 400 ºC. Zinc oxide absorbs

several different sulfur species such as hydrogen sulfides (H2S), carbonyl sulfide and

mercaptans. However, cyclic organic sulfur compounds such as thiophenes normally require

hydrogenation over Co-Mo or Ni-Mo catalysts to H2S. The H2S is then adsorbed over the

zinc-oxide bed. The hydrogenation and zinc-oxide catalysts can be combined into one vessel.

Residual sulfur content is < 0.2 mg/m3 [Ullman’s, 1985].

There are four types of burner configurations used in tubular reforming as shown in Figure 5.

Top-fired reformers can have several parallel rows of tubes, while wall- or side-fired

reformers only can have one row. The homogeneity of the heat transfer to the tubes is

determined by burner geometry, flame length and diameter, tube-to-tube and row-to-row

spacing, fired tube length and distance from the flame to the reformer wall. The temperature

of the flame is approximately 1800 C at the hottest place, and 1100 C at the coldest place.

The heat transfer in the wall-fired reformers is mainly by the radiant side-wall, while in top-

fired reformers, the heat is transferred through radiation from the flame and hot flue gases

[Ullman’s, 1985].

Page 14: Natural Gas Conversion

10

Figure 5. Typical configurations of reformer furnaces [Dybkjaer, 1995].

Typical reformer wall temperature profiles in the top- and side-fired reformers are shown in

Figure 6 (left). The top-fired reformer is characterized by a temperature peak in the top, and it

has the highest heat flux where the metal temperature is at its maximum. The side-fired

reformer allows a better temperature control, and the maximum temperature is at the outlet of

the tube. The highest heat flux is at a rather low temperature. The side-fired reformer has a

higher average heat flux than the top fired. Moreover, the short residence time in the flames in

the side fired reformer ensures very low emissions of NOx in the flue gases [Dybkjaer, 1995].

A central problem in steam reforming is to balance the heat input through the tube with the

heat consumption by the endothermic reforming reaction, while at the same time limiting the

stress on the tubes by minimising the maximum tube wall temperature and the maximum tube

wall temperature difference. The side-fired reformer is ideally suited to solve this complex

interrelation [www.topsoe.com]. Godfrey and Shackcloth [1970] and Stehlik [1991] have

made some modelling of the radiation heat transfer from burner to steam reformer tube.

Figure 6. Left: Tube wall temperature and heat flux profiles. Top- and side-fired reformers. Start of run. Right:

Tube wall temperature profiles at start and end of run. Top- and side-fired reformers [Dybkjaer, 1995]

Page 15: Natural Gas Conversion

11

The catalyst will deactivate and lose activity upon use. It is however actually possible to

retain the productivity essentially constant by a slight increase in temperature in the lower end

of the tube, provided the required heat is available. A decrease in activity will however result

in a large temperature increase in the top of the tube, especially in case of a top-fired reformer

(see Figure 6 (right)). This is since the reaction consumes less heat when the catalyst is

deactivated, and hence the driving force for heat transfer to the process gas decreases. Due to

these effects, the top fired reformers must be designed with a considerable margin above the

maximum temperature at the start of the run. There is also risk of carbon formation at the hot

wall [Dybkjaer, 1995].

In the side-fired reformer, a decrease in catalyst activity will lead to an increase in

temperature in the upper part, as well, but the temperature will still be highest in the lower

end. Therefore, in this case, the reformer does not have to be designed for much higher

temperatures than at the start of the run [Dybkjaer, 1995].

Carbon formation

As already mentioned above, a critical parameter in steam reforming is the molar S/C molar

ratio. Carbon deposits will occur that deactivates the catalyst by cooking, if this ratio is too

low. Large carbon deposits may also block the tubes and hence cause hot-spots, which might

destroy the tubes. A higher S/C ratio reduces the carbon formation, and a common

steam/carbon ratio lies between 2.5 and 4.5, the higher value being used for steam reforming

of naphtha. A higher S/C ratio than stoichiometric, also helps to shift the steam reforming (1)

equilibrium towards the products, and hence to increase the methane conversion (reduce the

methane slip) [Moulijn et al., 2003]. The carbon is formed according to [Ullman’s, 1985]:

2CO C + CO2 (The Boudouard reaction)

CH4 C + 2H2 (Decomposition of methane)

CO + H2 C + H2O (Heterogeneous water gas reaction)

The equilibrium for the Boudouard reaction and the decomposition of methane are shown in

Figure 7 and 8. However, it has been shown that for some catalysts the thermodynamically

predicted carbon formation can be suppressed at temperatures below 700 ºC. See for instance

line (f) in Figure 7, which shows the regions where carbon was formed and not for a nickel-

uranium catalyst, and compare with (d) or (e) that show the true thermodynamic constant for

the Boudouard reaction. Theoretically, carbon could form in the inlet of the steam reformer.

However, this does not occur below 650 ºC, but only at higher temperatures when the catalyst

has a low activity so that it fails to convert sufficient methane. The addition of CO2 to the feed

gas minimizes the risk of carbon formation as shown in Figure 7 and 8 [Ullman’s, 1985].

Page 16: Natural Gas Conversion

12

Figure 7. Reaction quote for Boudouard reaction (5): 2CO C + CO2 as a function of temperature. (a)

condition of rich gas from naphtha at inlet to tubular reformer, (b) equilibrium reached for WGS (2), (c)

equilibrium reached for SR of methane (1) and WGS (2), (d) and (e) show the equilibrium constant for the

Boudouard reaction according to two different references in Ullman’s [1985], (f) boarder between carbon

formation and no carbon formation for a nickel-uranium catalyst, (g) boarder between carbon formation and no

carbon formation for a sulfided catalyst. [Ullman’s, 1985]

Figure 8. Reaction quote for methane decomposition (4): CH4 C + 2H2 as a function of temperature. (a)

equilibrium reached for SR of methane (1) and WGS (2), (b) equilibrium reached for WGS (2), (c) equilibrium

constant for methane decomposition (4), (d) and (e) show working lines for methane reforming in a reactor tube

with a high activity catalyst and a low activity catalyst, respectively (figures at the lines are fractions of tube

length from reactor inlets) (f) equilibrium constant determined by experiments from a reference in Ullman’s

[1985], (g) connecting line between equilibrium constants determined by experiments at 410 and 525 ºC from

two references in Ullman’s [1985].

Page 17: Natural Gas Conversion

13

As described above, traditionally the S/C ratios have been rather high in order to avoid carbon

formation. However, new highly active steam reforming catalysts and the use of adiabatic pre-

reformers make it possible to use S/C ratios below 1.0. The lower S/C ratio is advantageous if

a CO-rich gas is wanted, and to make the overall reaction less endothermic [Dybkjaer, 1995].

In a pre-reformer (see 2.3 Adiabatic pre-reforming), the worst carbon precursors (i.e. higher

hydrocarbons) are removed up-stream the steam reformer. This is the reason for why the main

reformer is able to operate at a lower S/C ratio [Moulijn et al., 2003].

If CO2 is added to the reformer feed in order to make the syngas richer in CO, as discussed

above, less steam is actually needed for the suppression of carbon formation since CO2 is an

oxidizing agent [Moulijn et al., 2003].

2.3 Adiabatic pre-reforming

Adiabatic pre-reforming is used for reforming of natural gas to heavy naphtha. The process is

carried out in a fixed-bed upstream the tubular reformer (see Figure 9). Higher hydrocarbons

are completely converted into CO, H2 and CH4. The reactions taking place in the pre-

reformer, except for WGS, are [Dybkjaer, 1995]:

CnHm + nH2O nCO + (n+m/2)H2 (10)

3H2 + CO CH4 + H2O (reversed 1)

In case of natural gas as feedstock, the overall process is endothermic, and therefore will

result in a temperature drop since the reactor is adiabatic. For higher hydrocarbon feedstocks,

the pre-reforming is exothermic or thermoneutral. In the pre-reformer, the temperature is

relatively low. This makes the chemisorption of sulfur to the Ni-catalyst favorable. Therefore,

traces of sulfur from the desulfurization unit will be trapped in the pre-reformer [Dybkjaer,

1995].

Figure 9. Installation of a pre-reformer [Dybkjaer, 1995].

Advantages of installing a pre-reformer [Dybkjaer, 1995]:

Page 18: Natural Gas Conversion

14

- All higher hydrocarbons are completely converted into CO, H2 and CH4. Even for

naphtha, the natural gas steam reforming catalyst may be used in the prereformer.

- All traces of sulfur are removed, which increases the life-time of the tubular steam

reforming catalyst.

- Since no sulfur is poisoning the top layer of the catalyst in the steam reformer, risks of

hot spots are reduced.

- The production capacity of the plant may be increased. By installing a new pre-heat

coil or heater between the pre-reformer and the steam reformer, the load on the

reformer is reduced. This may be used as a capacity increase or as a decrease in firing

with unchanged capacity.

- The sensitivity of the steam reformer to variations in S/C ratios and feedstock

composition is essentially eliminated.

2.4 Partial oxidation

Partial oxidation (POX) is often used for gasification of heavy oil, but all hydrocarbons are

possible as feedstocks. Since no water is added, the H2/CO ratio is lower than compared with

steam reforming or autothermal reforming. Partial oxidation of natural gas is used in small

plants and in regions where natural gas is cheap. Reaction temperatures are 1350 – 1600 C

and pressures up to 150 bar. The concentrations of different compounds in the product

mixture are determined by several equilibria, which are quickly tuned in at these high

temperatures. Partial oxidation can be performed with or without a catalyst. If a catalyst is

used, the reaction temperature can be lower, the reactions still reaching equilibrium, since the

catalyst lowers the activation energies. A lower temperature, at this high temperatures, would

not significantly change the equilibrium composition [Ullman’s, 1985].

Bharadwaj and Schmidt [1995] claimed that the development over the years has been to

minimize the use of steam in reforming due to the following disadvantages:

- Endothermic reactions

- The product gas has a H2/CO ratio of 3

- Steam corrosion problems

- Costs in handling excess H2O

The trend is to move from steam reforming to “wet” oxidation (autothermal reforming) to

“dry” oxidation. The dry oxidation is the partial oxidation of methane (6). This reaction

directly gives the desired ratio H2/CO = 2 for Fischer-Tropsch or methanol synthesis, at

temperatures > 900 ºC, as shown in Figure 1 (b) above. However, the selectivites are also

affected by the H2O and CO2 formed in the complete oxidation reactions (7, 10) [Bharadwaj

and Schmidt, 1995].

Since the partial oxidation reaction is slightly exothermic, the partial oxidation reactor would

be much more energy efficient than the energy intensive steam reformer. Since the reaction is

fast, the reactor size could be reduced significantly compared to the SR reactor [Bharadwaj

and Schmidt, 1995].

The non-catalytic partial oxidation (POX) of methane is already commercialized, for instance

in the Fischer-Tropsch plant in Malaysia, Bintulu. However, there are several problems with

POX that can be avoided by catalytic partial oxidation (CPOX). In CPOX, only

heterogeneous reactions are taking place, lower temperatures are used and no soot or

unwanted by-products are formed. No burner is used. It is mainly the absence of

Page 19: Natural Gas Conversion

15

homogeneous reactions that prevent the formation of unwanted oxidation products or flames

which can lead to soot formation. The newly developed highly selective partial oxidation

catalysts can overcome the carbon problem without any steam input. It appears that high

selectivity to CO and H2, as compared to CO2 and H2O, can be achieved at short contact

times, i.e. at high GHSVs. Results showing high methane conversion at short contact times

have also been reported. Bharadwaj and Schmidt [1995] concluded that for partial oxidation

of methane, Rh seems to be the catalyst of choice to achieve selectivities to H2 and CO above

95 %, and methane conversions above 90 %. The major engineering challenge is to ensure

safe operation with the premixed CH4/O2 mixtures. The reaction mechanism seemed to be

direct oxidation via CH4 pyrolysis at contact times shorter than 0.1 s. At longer residence

times, reforming reactions with CH4 and H2O or CO2 and shift reactions also may take place.

However, the direct catalytic partial oxidation of methane has not yet been commercialized

because it is difficult to study since it involves premixing of CH4/O2 mixtures which can be

flammable or explosive [Bharadwaj and Schmidt, 1995].

Due to the high temperatures needed in order to reach high conversions and high selectivities

to H2 and CO, the CPOX-reactor needs extremely tolerant materials, which are expensive.

Current research at NTNU/SINTEF aims at developing catalysts that can give satisfactory

high reaction rates and high selectivities to H2 and CO at lower temperatures (650 ºC). This

would be possible if the catalyst could ensure that only the partial oxidation, i.e. reaction (6),

occurs, and hence that no water is formed. Reaction (6) is exothermic and hence favored at

low temperatures, and if no water is ever present in the reactor, a syngas with a H2/CO ratio of

2.0 could theoretically be formed at low temperatures in relatively cheap steel reactors

[Bjørgum, 2006].

According to Dybkjaer [1995] and Lødeng [2006], this far, no catalyst has shown selectivity

for CPOX of methane to CO and H2 at industrial conditions (i.e. at pressures above 20 bar)

yielding less CO2 than predicted by equilibrium of methane steam reforming and WGS. In

order to be economically attractive, catalytic POX of methane must be carried out at elevated

pressure and at high methane conversion. The composition of the product gas will always be

determined by the equilibrium at the reactor exit, irrespective of whether a fixed or fluidized

bed is used [Dybkjaer, 1995].

2.5 Autothermal reforming and secondary reforming

In autothermal reforming (ATR), partial oxidation of the fuel is conducted in order to produce

the heat required for the endothermic reforming reactions of the same fuel. External steam is

added. Characteristic of the autothermal process is that the oxygen added to generate heat is

chemically bound in the product gas, which results in that the H2/CO ratio in the product gas

is lower than in other processes. Both temperature and pressure is in between what is used in

steam reforming and partial oxidation. Typical operating conditions are 850 – 1000 C and 20

– 100 bar [Moulijn et al., 2003]. One advantage with autothermal reforming is that the

thermal efficiency (i.e. ratio of heat content of reformed gas to that of the hydrocarbon feed, at

0 C) is higher (88.5 %) than that of steam reforming (81 %, including fuel) and than that of

partial oxidation (83.5 %). The maximum temperature is not limited by the tube material but

by the stability of the catalyst and the refractory lining of the reactor. Autothermal reforming

is more flexible than tubular reforming, since the higher operating temperature can

compensate for the increase in methane slip which higher pressure would cause otherwise

[Ullman’s, 1985].

Page 20: Natural Gas Conversion

16

There are two types of autothermal reformers – pure catalytic, and with a pre-combustion

unit. In the pure catalytic reformer, the reactants (methane, oxygen and steam) enter the

catalyst bed directly after mixing. This performance, without any residence time in the empty

space above the catalyst bed, results in no carbon formation, even at low preheat temperatures

of the feedstock (methane). Steam reforming reactions take place more quickly than the

Boudouard reaction equilibrium is tuned in. This makes it possible to use less steam, and

hence less oxygen, which results in a higher CO/CO2 ratio in the product gas. The drawbacks

with this purely catalytic process are high thermal and mechanical loads on the catalyst in the

immediate vicinity of the burner. Temperature variations during start-up and shut-down, and

high gas velocities lead to attrition and catalyst disintegration, which in turn makes it

necessary to replace the catalyst every two years [Ullman’s, 1985].

Reforming with pre-combustion in the empty space above the catalyst bed is preferred for

gases that have very low risk of carbon deposition, i.e. gases with high H2 content and low

hydrocarbon content. This type of reforming requires slightly more oxygen than the pure

catalytic type, and the gas velocities must be lower [Ullman’s, 1985]. For natural gas

autothermal reforming, the concept with the pre-combustion zone is preferred, and will be

described further.

ATR, as a concept, is a stand-alone process in which the entire natural gas conversion is

carried out by means of internal combustion with oxygen. Secondary reforming is a process in

which partially converted process gas from a tubular steam reformer is further converted by

means of internal combustion. The secondary reformer in an ammonia plant will be air-blown,

while that of a methanol plant will be oxygen blown. [Dybkjaer, 1995]

The autothermal reforming is usually not used on its own, due to high investment costs for

separation of the oxygen from air. It is rather used downstream a steam reformer, i.e. as a

secondary reformer, in order to reform the unreacted methane from the steam reformer. This

makes it possible to increase the pressure, without increasing the temperature, in the steam

reformer (typical conditions are 1070 K and 30 bar), which is not favourable for the

equilibrium but economically favourable if the syngas will be used for a high-pressure

chemical synthesis. In case of ammonia production (> 100 bar), N2 is needed in the synthesis

and hence air can be used instead of oxygen in the autothermal reformer, which reduces the

cost for this reformer considerably. The two-step reforming concept with a SR combined with

a secondary reformer (ATR) is suitable also for methanol syngas production, since the

methanol takes place at high pressures (50 – 100 bar) [Moulijn et al., 2003].

Since the concentration of combustibles are different in the gas going to the ATR and the

secondary reformer, different designs of burner and reactor are needed, in order to ensure

suitable heat release and avoid soot formation in each case. The ATR and the secondary

reformer are however very similar in reactor design. They consist of a refractory-lined

pressure vessel (and therefore stands higher pressures and temperatures than the steam

reformer) with a burner, combustion chamber and catalytic bed. The reactor space can be

divided into three zones (see Figure 10) in which different reactions take place according to

Table 3. [Dybkjaer, 1995]

Page 21: Natural Gas Conversion

17

Figure 10. Schematic picture of an ATR [Moulijn et al., 2003].

Table 3. Reaction zones in an ATR

Reactor equipment Reaction zones

Burner - (provide mixing)

Combustion

chamber

Combustion zone

Thermal zone

Catalyst bed Catalytic zone

Burner:

The burner provides the mixing of the feedstreams in a turbulent diffusion flame. The core of

the flame has a high temperature, often above 2000 ºC, why it is important to minimize the

transfer of heat back to the burner from the flame. This can be made by recirculating the gas

from the thermal zone back to the burner [Dybkjaer, 1995]. The burner is the key element for

the oxygen-fired reformer. Careful design of the burner nozzles ensures a flow pattern with

efficient mixing that protects the refractory and burner from the hot flame core

[www.topsoe.com]

Combustion zone:

This is the turbulent zone in which the hydrocarbon and oxygen are mixed and combusted.

Often the principle “mixed-is burnt” is valid, since the exothermic combustion reactions are

very fast. The overall oxygen to hydrocarbon ratios in the combustion zone vary between 0.55

and 0.6, which means that the conditions are substoichiometric with respect to complete

combustion. The combustion reactions are numerous complex radical reactions, but for

modeling purposes, it is often enough to describe the reactions by an overall one molecular

reaction. In case of natural gas [Dybkjaer, 1995]:

CH4 + 3/2 O2 CO + 2H2O (11)

This reaction is a mixture of (6) and (7). All oxygen is consumed in the combustion zone, and

the unconverted CH4 will continue down to the thermal zone.

Page 22: Natural Gas Conversion

18

In case of a secondary reformer, also H2 will be burnt to water in the combustion zone

according to reaction (9).

Thermal zone:

In the thermal zone, the conversion of the hydrocarbon proceeds via homogeneous gas phase

reactions. The main reactions are thermal methane reforming (1) and WGS (2). Also pyrolysis

of higher hydrocarbons takes place. [Dybkjaer, 1995]

Catalytic zone:

The catalytic zone is a fixed-bed, in which the hydrocarbons are finally converted through

heterogeneous catalytic reactions. At the exit of the catalytic zone, the gas mixture will be in

equilibrium with respect to reactions (1) and (2) at the exit temperature and pressure, and the

catalyst will destroy any soot precursors formed in the combustion chamber. The syngas is

completely free of oxygen. The top of the catalyst bed is exposed to gases with temperatures

of 1100 – 1400 ºC. A Ni-catalyst supported on a magnesia-alumina spinel has the required

activity and high-temperature stability. The overall reaction rate is mainly controlled by the

external diffusion rate, i.e. the transport rate of the reactants through the gas film surrounding

the catalyst pellets. This means that the process can be carried out at very high space

velocities, since the catalytic reaction is very fast. It is also known that higher space velocities

reduce the film thickness surrounding the pellets. The amount of catalyst needed is actually

determined by optimal flow distribution and pressure drop in the reactor [Dybkjaer, 1995]. In

synthesis gas production, a S/C ratio as low as 0.6 is industrially proven. Soot-free operation

is achieved through optimised burner design and by catalytic conversion of soot precursors

over the catalyst bed [www.topsoe.com]. The ATR or secondary reformer is operated close to

adiabatically, and hence the temperature is given by the adiabatic heat balance [Aasberg-

Petersen et al., 2003].

Applications of the concepts ATR and secondary reforming are discussed in 3. Different

reforming concepts for different applications.

When steam reforming and autothermal reforming are combined, it is logical to use the heat

produced in the latter to warm the tubes in the former one. This configuration is called heat-

exchange reformer, and is described below.

2.6 Heat exchange reforming

The idea with heat exchange reforming is that part of the heat is supplied to the tubes by heat

exchanging with process gas. Heat exchange reforming eliminates the expensive fired

reformer. In this configuration, only medium pressure steam can be recovered from the syngas

plant and electricity for the syngas compressor must be imported. Figure 11 shows two

different types of heat-exchange reformer. The left is the ICI combined reforming process and

the right the Exxon CAR (combined autothermal reforming) process. In the ICI process, 75 %

of the methane is reformed in the steam reformer at 970 K and 40 bar, and the rest is

converted in the autothermal reformer. In the Exxon CAR process, steam reforming and

partial oxidation is combined in a single reactor and reaction zone. The reactor is a fluidised

bed, and oxygen is introduced in the middle of the bed [Moulijn et al., 2003].

Page 23: Natural Gas Conversion

19

Figure 11. ICI combined reforming process (left) and Exxon CAR process (right)

In the case where two reformers are combined, the heat needed in the tubular SR reformer is

obtained from the hot product gas from the second reformer. The second reformer can be an

air- or oxygen fired ATR or a partial oxidation unit. This concept is used for production of

hydrogen or syngas for the methanol synthesis. In case of syngas production for the ammonia

synthesis, there is a poor correspondence between the heat required and the heat available in

the air-fired secondary reformer, and hence the concept of heat exchange reformer is not

common [Dybkjaer, 1995].

A problem associated with the heat exchange reforming described above, in which CO-rich

gases contact metals at high temperatures, is the risk of metal dusting corrosion. A certain gas

mixture will potentially form carbon via the exothermic Boudouard reaction at temperatures

below which the mixture satisfies the Boudouard reaction equilibrium. The temperature below

which a certain gas mixture may form carbon is called the Boudouard temperature. For a gas

mixture with a high Boudouard temperature, i.e. a CO-rich gas, the Boudouard reaction may

be catalyzed by hot metal surfaces. Therefore, it is important that a gas mixture with a high

Boudouard temperature does not come in contact with a metal surface of a slightly lower

temperature. If carbon deposition on the metal starts, there is a big risk of metal corrosion. If

carbon is deposited on the catalyst, this will of course lead to deactivation. Therefore, the

catalyst outlet temperature is always higher than the Boudouard temperature. However, a

temperature drop in the reformer could make carbon deposition possible [Dybkjaer, 1995].

Another type of heat exchange reformer is the convection reformer. In this reformer, the tubes

are heated mainly by the flue gas flowing upwards on the outside of the tubes, but also by the

reformer gas flowing upwards inside the tube [Dybkjaer, 1995].

The Topsoe convection reformer, see Figure 12 and 13, has a single burner, which simplifies

the design and control of the unit. The burner is separated from the tube section. It is called

convection reformer since it essentially means that the radiant tube section and the hot part of

the convection section are combined in a relatively small unit. The exit temperature from the

reformer is approximately 600 ºC for both product gas and flue gas after heat exchange. This

reduction in temperature means that 80 % of the fired duty is utilized in the process, compared

to 50 % in a conventional SR. In plants using the convection reformer, it is possible to

Page 24: Natural Gas Conversion

20

balance the energy need of the reformer by the energy available in the PSA off-gas, hence

avoiding the energy surplus in conventional systems [Dybkjaer, 1995].

Figure 12. The Haldor Topsoe Convection Reformer, HTCR, for hydrogen production [www.topsoe.com].

Figure 13. HTCR reformer (left) with burner, principle of heat transfer in the tubes (right) [www.topsoe.com]

Page 25: Natural Gas Conversion

21

3. Different reforming concepts for different applications

3.1 Introduction

A typical ATR plant is shown in Figure 14. It comprises a feed preheat section, an ATR-

reactor, a heat recovery section and a gas separation unit. The ATR plant contains much less

equipment than the conventional SR plant. Desulfurization is not needed if the feed is low-

sulfur natural gas. The schematic picture in Figure 14 shows the lay-out of an ATR plant for

the production of a syngas of H2/CO-ratio of 2.0. Such a lay-out can be used for syngas

production to a methanol plant, Fischer-Tropsch plant, for hydrogen production or for syngas

to an ammonia plant. In the ammonia case, air has to be used in the ATR [Dybkjaer, 1995].

Figure 14. Typical process lay-out for an ATR, for production of a syngas with H2/CO ratio of 2.0 [Dybkjaer,

1995].

The concept with secondary reforming is often used for syngas production to methanol or

ammonia synthesis. Figure 15 shows the lay-out of a plant for production of ammonia syngas.

Except from the tubular SR reformer and the secondary air-blown reformer, Figure 15 also

contains a pre-reformer. In case of natural gas as feedstock, the pre-reformer is usually not

necessary in the production of ammonia syngas. The concept with the pre-reformer is only

interesting when the steam production must be minimized.

Page 26: Natural Gas Conversion

22

Figure 15. State-of-the-art plant for production of ammonia synthesis gas [Dybkjaer, 1995].

Figure 16 shows the lay-out of the two-step production of syngas for methanol synthesis. The

conditions in the primary SR unit are mild. In the second reformer, oxygen and steam are

used. [Dybkjaer, 1995]

Figure 16. Two-step reforming plant for production of methanol synthesis gas [Dybkjaer, 1995].

The methanol synthesis takes place at 50 – 100 bar, and hence also for this application a

steam reformer and an autothermal reformer in series is sometimes used. However, since pure

oxygen is needed for the methanol synthesis, the autothermal reformer will be rather

expensive. Therefore, more than 80 % of the syngas production for methanol synthesis is

made solely with steam reforming of methane [Moulijn et al., 2003].

Page 27: Natural Gas Conversion

23

3.2 Methanol synthesis

The ideal syngas composition for methanol production is H2/CO = 2 and 5 vol% of CO2,

which increases the activity. It is actually the CO2, and not the CO, that reacts with H2 on the

catalyst surface, Cu/ZnO/Al2O3, to give methanol.

CO + 2H2 CH3OH ΔHR,298K = -90.8 kJ/mol

CO2 + 3H2 CH3OH + H2O ΔHR,298K = -49.6 kJ/mol

CO + H2O CO2 + H2 ΔHR,298K = -41 kJ/mol

The newly developed methanol catalysts have selectivities of up to 99 %. This is essential

since methanol is thermodynamically less stable than other possible products, for instance

CH4. The methanol synthesis is limited by equilibrium, why it is important to keep the

temperature low. This is obtained by quenching the reaction by cold feedgas at different

places in the bed as in the ICI-process (see Figure 17), or by several reactors with

intermediate cooling as in Haldor Topsoes process. Thus, the catalysts must be active at low

temperatures. Recycling is always performed in order to increase the overall yield. Typical

conditions for the methanol synthesis today are 50 – 100 bar and 500 – 550 K. Steam

reforming of methane is the most common way to produce the syngas, in which a H2/CO ratio

of 3 is obtained [Moulijn et al., 2003].

Figure 17. Flow scheme of the ICI low-pressure methanol plant [Moulijn et al., 2003].

In order to correct the H2/CO/CO2 ratio, prior to the methanol synthesis, either the surplus of

hydrogen is burnt or the CO content is increased. This can be done in two ways:

1. If CO2 is available it may be added to the reformer or to the raw syngas.

2. By installing an oxygen-fired ATR downstream of the SR, or by only using an ATR with

adjustment by CO2 removal in order to correct the carbon oxide content.

Page 28: Natural Gas Conversion

24

For both SR and ATR, the exit gas mixture is in equilibrium.

HP-steam is often produced at the exit of the ATR, or from the effluent gases from the

burners in the SR. This HP-steam can for instance be used to drive syngas compressors. When

SR and ATR are combined and heat exchanged, the high-temperature heat in the product gas

from the ATR is used to heat the steam reformer tubes, and hence only MP-steam can be

produced in this case.

3.3 Ammonia synthesis

The syngas for ammonia production requires a H2/N2 ratio of 3, which in 80 % is made from

steam reforming of natural gas followed by ATR with air.

N2 + 3H2 2NH3 ΔHR,298K = -91.44 kJ/mol

The equilibrium is favoured by low temperature and high pressure. However, the reaction rate

is very low below 670 K. The typical ammonia synthesis conditions are 675 K at inlet, 720 –

770 K at exit, and 100 – 250 bar. The same reactor solutions as used in methanol synthesis are

used in the ammonia synthesis, hence cold-feed quenching (see Figure 18) or several reactors

with intermediate cooling. The single-pass conversion is low (20 – 30 %) since the reaction is

limited by equilibrium, and hence recycling is necessary [Moulijn et al., 2003].

.

Figure 18. ICI quench reactor and temperature-concentration profile [Moulijn et al., 2003]

3.4 Fischer-Tropsch synthesis

The fundamentals in Fischer-Tropsch (FT) synthesis have been described thoroughly by

Lögdberg [2006]. Here, the choice of syngas production concept will be discussed. Figure 19

gives a schematic FT lay-out.

The preparation of synthesis gas from natural gas stands for 50 – 75 % of the capital cost in a

Fischer-Tropsch plant. According to Aasberg-Petersen et al. [2003], oxygen-blown ATR is

considered to be the best option for large-scale, safe and economic syngas production. Air-

blown ATR has shown to be less efficient and more cost intensive considering the whole

Page 29: Natural Gas Conversion

25

GTL-plant. The energy consumption for the air-blown alternative is much higher. The energy

consumption in the syngas section alone is 10 % higher for the air-blown ATR compared to

the oxygen-blown. The natural gas is often pre-reformed in an adiabatic reactor in order to

convert higher hydrocarbons to CH4, H2 and CO, by steam addition. The use of a pre-reformer

reduces the overall oxygen consumption.

According to Wilhelm et al. [2001], the syngas production for large-scale GTL plants in the

future will be the two-step reforming concept, or ultimately the ATR concept. They claimed

that air-blown reformers will never be competitive to oxygen-blown ones in GTL-plants,

although the investment cost for the air separation unit is eliminated. The reason is that the

air-blown reformer appears much less flexible, has lower thermal efficiency and requires high

air compression power. Furthermore, the usage of an air-blown reformer makes it impossible

to recirculate FT tail-gas, and the gas flows out from the reformer will be much larger which

in turn will increase the size of the equipment downstream the reformer. The FT-tailgas can

usually be used for power production by combustion in gas turbines. However, if it contains a

high concentration of N2, it will have a low heating value and the power production might not

be viable.

Figure 19. Schematic layout of Fischer-Tropsch plant [Aasberg-Petersen et al., 2003].

Steam to carbon ratios as low as 0.6 inside the ATR have been proven industrially. The

desired H2/CO ratio out from the ATR is, in case of FT, 2.0. This low ratio can only be

achieved at extremely low S/C ratios, see Figure 20, without recirculating CO2 [Aasberg-

Petersen et al., 2003].

Figure 20. H2/CO ratio as a function of S/C ratio in ATR outlet gas [Aasberg-Petersen et al., 2003].

Page 30: Natural Gas Conversion

26

Operation at low S/C ratios improves process economics, which is shown in Table 4.

However, low S/C increases the risk of carbon formation and hence catalyst deactivation. A

lot of research is going on in order to elucidate the causes of carbon formation, and the means

to prevent carbon formation or to destroy the carbon in the catalytic bed. Carbon formation

has been found to depend upon feed gas composition, temperature, pressure and especially

burner design. Pilot plant demonstrations have succeeded to run soot free operations at S/C

ratios as low as 0.2 [Aasberg-Petersen et al., 2003].

Table 4. Process economics with different S/C ratios [Aasberg-Petersen et al., 2003].

Bharadwaj and Schmidt [1995] seemed to have another opinion than Asberg-Petersen et al.

[2003] and Wilhelm et al. [2001]. They claimed that partial oxidation is much faster, highly

selective in a single reactor, and much more energy efficient than SR and ATR. It could hence

reduce the capital and operating costs of syngas production, and would hence be the choice of

reforming concept in the future. However, the catalytic partial oxidation (CPOX) is still not

commercialized, and was discussed more in 2.4 Partial oxidation.

According to Aasberg-Petersen et al. [2003], CPOX suffers from a highly flammable mixture

upstream the reactor at conditions relevant to GTL. The autoignition temperature is

approximately 250 ºC for the mixture, therefore the inlet temperature of the mixture to the

reactor is low, typically 200 ºC. Compared to an ATR with a pre-reformer, the CPOX concept

consumes more natural gas and oxygen for the same productivity. This is due to that much of

the natural gas must be completely oxidized in order to release enough energy to raise the

temperature in the reactor to the desired exit temperature. The exit compositions will be

essentially the same in both concepts, provided the amount and properties of the inlet streams

are the same. The authors meant that CPOX is not economical especially since it requires a lot

more oxygen than the ATR concept. Furthermore, the inherent safety constraints make the

CPOX less attractive. Air-blown CPOX could be an alternative in case of making fuels for

fuel cells, however not for FT or methanol synthesis, as discussed above.

4. New reforming concepts under development

Normal steam reforming accounts for 60 – 70 % of the investment costs of a methanol or FT-

plant based on natural gas, and the air separation unit stands for a major part. The

combination of steam reforming and autothermal reforming (the heat exchange reformer

concept) could to some extent reduce the investment and operating costs, and especially the

catalytic partial oxidation (CPOX) has a large potential to reduce the syngas production cost,

according to Moulijn et al. [2003]. CPOX is attractive since it has a high exergy efficiency,

which is due to the fact that no heat exchange is needed between a hot side and a cold side

since all reactions take place in the same vessel. Also the ICI combined reforming process

gives approximately 30 % lower exergy losses compared to a conventional steam reformer.

This is mainly due to that the need for irreversible combustion reactions to heat the furnace in

the steam reformer case is eliminated in the combined reforming process. However, Moulijn

et al. [2003] claimed that “a major breakthrough would be required to make any substantial

Page 31: Natural Gas Conversion

27

savings in the cost of producing syngas”. Some examples of such new ideas, still on the

development/demonstration stage are discussed below.

As discussed above, the air-separation unit is a cost intensive part of the syngas preparation in

the methanol and FT cases. It has, however, been shown economically advantageous to run

the ATR-burner on oxygen compared to air, as discussed above. A more promising alternative

to eliminate the air-separation unit would be a ceramic membrane reactor. In this concept, an

ionic or oxygen transporting membrane makes it possible to combine air separation and

partial oxidation in one unit. The development of such membranes is still relatively early. The

membranes typically work at 750 ºC or higher. Oxygen is transported through the membrane

with a selectivity of 100 %. On the inside of the membrane it reacts with the hydrocarbon. If

the air must be compressed in order to ensure similar pressures on the two sides, the concept

might not be economical. If ambient pressure air can be used, the membrane must have a high

mechanical strength in order not to break at the high inner pressures desired for the syngas

production. The concept with ceramic membrane reforming seems promising, but its

feasibility must be proven [Aasberg-Petersen et al., 2003].

Membranes can also be used in order to remove the produced hydrogen from the product gas

(see Figure 21). This is called membrane reforming. In this way, the amount of hydrogen

produced is not limited by the equilibrium and hence much lower temperatures can be used,

and no traditional CO2 removal system is needed. [Moulijn et al., 2003]

Figure 21. Schematic picture of membrane reforming [www.co2captureproject.org].

Another technique under development is the chemical looping reforming, in which chemically

bound oxygen is used to oxidize methane. The system consists of two reactors, as shown in

Figure 22. In the air-reactor, metal particles are oxidized to metal oxide by air (exothermic

reaction), and in the fuel reactor, methane is partially oxidized to syngas by means of the

oxygen in the metal oxide (endothermic reaction). The oxygen carrier is continuously

circulated between the two beds, thus exchanging both oxygen and heat

[www.co2captureproject.org]. A possible alternative would be to install steam reformer tubes

in the fuel reactor, letting the complete oxidation of methane on the fluidized metal oxide

particles give heat to the steam reforming reactions inside of the tubes. Actually, the heat used

for the steam reforming would be released when the oxygen carrier is oxidized in the air

reactor, and transported to the fuel reactor with the metal oxide particles. The complete

combustion of methane outside of the steam reformer tubes is actually endothermic and this

process is needed in order to reduce the metal oxide back to the metal [Rydén, 2006]. The

advantage with chemical looping reforming is that the separation of oxygen from air is simple

and, in case of complete combustion, the CO2 can be captured easily since it is free from N2.

Another advantage is that partial oxidation could be performed without forming the explosive

gas mixtures of oxygen and methane discussed in 2.4 Partial oxidation.

Page 32: Natural Gas Conversion

28

Figure 22. Schematic picture of chemical looping partial oxidation [www.co2captureproject.org].

Another concept that implies CO2 capture from hot flue gases, which is gaining interest due to

the possibility of CO2 sequestration, combined with hydrogen production, is the sorption

enhanced steam reforming process (SERP). The idea is to mix the ordinary steam reforming

catalyst with a CO2 adsorbent, e.g. dolomite or lithium zirconate. CO2 is captured, which

shifts the equilibriums of reactions (1) and (2) towards hydrogen, and makes it possible to

perform the steam reforming reaction at much lower temperatures (450 – 630 ºC) than

conventional. Hence, investment and operational costs may be significantly reduced. It is

possible to obtain a product gas containing 97 % H2 on a dry basis. After saturation, the CO2-

saturated adsorbent is regenerated by heating. Kinetic limitations of the adsorbent however

hinder the concept from being commercialized [Ochoa Fernandéz et al., 2005].

5. Some notes on modeling of CPOX and ATR

De Groote et al. [1996] made a one-dimensional heterogeneous model to simulate the

adiabatic CPOX of methane in a fixed-bed reactor. A plug-flow was assumed. The reactions

considered were CO2 reforming, steam reforming, WGS, complete combustion, the

Boudouard reaction and methane cracking, for which kinetic rate expressions were set up.

Internal gradients in the catalyst pellets were accounted for by using effectiveness factors.

These effectiveness factors were very small (between 0.05 and 0.07) for all reactions except

for the WGS (0.7). It was concluded that when air is used as oxidant, or when steam and CO2

are added to the feed, acceptable temperatures are achieved in the bed. When oxygen is used,

Page 33: Natural Gas Conversion

29

a temperature maximum of 1500 ºC was achieved, which would destroy the Ni-catalyst.

When steam is added, the amount of coke becomes negligible. CO2 also reduces carbon

formation but increases the area in which coke is formed.

Hoang and Chan [2004] made a two-dimensional heterogeneous reactor model for an

adiabatic ATR. The application of the ATR was for hydrogen production to fuel cells. They

only included four reactions, i.e. the complete combustion of methane, methane steam

reforming to CO and H2 (and to CO2) and WGS. The carbon deposition reactions were

neglected since the chosen S/C and air to carbon (A/C) ratios would ensure no formation of

carbon. The combustion chamber was actually not modeled, but the complete reactor was

assumed filled with Ni-catalyst. It was assumed that the inlet temperature of the catalyst bed

was approximately 300 ºC, which is the light-off temperature, i.e. the temperature at which

the autothermal reaction can be self-activated.

Biesheuvel et al. [2003] made a one-dimensional ATR-reactor model for conversion of

methane. They divided the model in two sections: an upstream oxidation section and a

downstream reforming section. In the oxidation section, all of the oxygen is converted, with

partial conversion of the fuel. An empirical fuel utilization ratio is used to quantify which part

of the feed is converted in the oxidation section as a function of the relative flows of air and

steam. In the oxidation section, the gas temperature rapidly increases toward the top

temperature at the intersection with the reforming section. In this section the temperature

decreases while the fuel is further converted with water and CO2 as oxidant.

For the oxidation section, no kinetics were considered. The (partial) oxidation reaction in this

zone was assumed to take place instantaneously, and all oxygen reacted, and hence no need

for a catalyst. Experimental data of the top-temperature (i.e. the temperature at the start of the

catalytic bed (the reforming section)) as a function of inlet flows led to an empirical

expression for the fuel utilization ratio, i.e. how much of the steam and CH4 that was

converted in the oxidation section and the selectivity to H2 and CO, and H2O and CO2. In

order to find this relation, the heat release and heat loss in the oxidation section and the Cp-

values of the gas mixtures must of course be taken into account. Probably, the composition of

the gas exiting the oxidation section was found by iteration by solving mass and energy

balances over the oxidation section, for a guessed outlet composition, until the calculated

temperature agreed with the measured. Hence the authors could correlate a measured top

temperature to a certain gas composition/flow at the inlet of the reforming section, and then

the actual modelling of the reformer section was made based on these inlet data.

Figure 23 and 24 show some results from the simulation.

Page 34: Natural Gas Conversion

30

Figure 23. Oxidation section in ATR. γ = S/C ratio. Top temperature on left y-axis. Utilization factor (dashed

lines, fraction of methane converted in the oxidation section) and thermodynamic CH4 conversion (solid lines)

on right y-axis. T0 = 400 ºC, adiabatic [Biesheuvel et al., 2003].

Figure 24. Temperature profiles in the reforming section at different S/C ratios. The exit temperature and the

inlet gas composition will describe the composition of the product gas. Oxygen/carbon ratio = 0.57 [Biesheuvel

et al., 2003].

Page 35: Natural Gas Conversion

31

Modeling of trickle-bed reactors

- The Fischer-Tropsch reaction

Page 36: Natural Gas Conversion

32

1. Fixed-bed reactors vs trickle-bed reactors

In a conventional fixed-bed reactor, only gases and solid phases (the catalyst particles) are

present. This makes the modeling rather straight-forward. However, if a liquid phase is added

to the fixed bed, the modelling becomes a bit more complex. This concept is called trickle-

bed, and often means that the reaction takes place between a gaseous and a liquid reactant, the

gaseous reactant having to diffuse through the liquid phase to the catalyst surface and there

react with the other reactant already present in the liquid.

The downflow cocurrent column packed with catalyst may operate in mainly two different

flow regimes: “the trickle flow” in which the gas phase is continuous and the liquid phase

dispersed, and “the bubble flow” in which the gas phase is dispersed and the liquid phase

continuous. Increasing the gas flow rate will eventually lead to pulsed flow. Figure 1

illustrates the different regimes.

Figure 1. Flow regimes in cocurrently operated packed bed. G is the gas load in [kg/m

2, s] and L the liquid load

in [kg/m2, s] [Gianetto and Silveston, 1986].

The flow patterns in a trickle-bed range between gas-continuous and liquid-continuous

regimes, depending on shape and size of particles as well as on physical properties and flow

velocities of the gas and liquid

Trickle-beds are widely used in the refineries for processes such as hydrodesulfurization of

different petroleum fractions, hydrocracking of heavy gasoils and atmospheric residues,

hydrotreating of lubricant oils and for hydrogenation [Froment and Bischoff, 1990]. Trickle-

beds are also often used in the Fischer-Tropsch synthesis, in which syngas is converted into

hydrocarbons. In this case, the reactants are gaseous and the products gases and liquids.

Page 37: Natural Gas Conversion

33

The trickle-bed, with its stationary catalyst bed, is preferred to a slurry type operation, in

which the catalyst is suspended in the liquid, when the gas flow rate is relatively low because

it leads to a gas and liquid flow pattern that better approximates the plug flow. A plug flow,

with its high reactant concentration, is favourable if the total reaction order is positive. For

high gas flows, the slurry operation might be better suited since it avoids the pulsed flow

regime that can come up in a fixed-bed. However, Huang et al. [2004] reported that there

might be some advantages to work in the pulsing flow mode. The heat transfer is often poor in

the conventional trickle-flow regime. This is undesirable especially if the reaction is

exothermic, due to risk of run-away, and also due to formation of undesired by-products. One

way to enhance the heat transfer in trickle-beds is to operate in a high gas-liquid interaction

regime called pulsing flow by Huang et al. [2004]. In the pulsing flow mode, flows of gas-

continuous bases and liquid-rich slugs (pulses) are alternated. The heat transfer rates in the

pulses can be 3 – 4 times higher that those in the gas-continuous bases. The pulsing flow can

also enhance the wetting of the catalyst pellets and hence improve the liquid distribution,

which in turn will minimize the risk for hot spot formation. In the pulsing flow mode also the

mass transfer rates and the liquid hold-up can be significantly higher than in the conventional

trickle-flow mode.

The main difference between the modeling of a fixed-bed and a trickle-bed is the

hydrodynamics because of two fluid phases in the latter case. Furthermore, due to the extra

liquid phase in the trickle-bed, more mass and heat transfer resistances have to be

encountered, end hence gas-liquid and liquid-solid interfacial areas must be known. For a

conventional fixed-bed, the mass transfer coefficient, kg, and the geometrical external particle

surface area, av, is enough. For the mass transfer over the gas/liquid interface, usually Henry’s

constant, H, is used to estimate the concentration of a component at the interface.

Figure 2 shows schematic concentration profiles of reactants and products in the different

phases existing in a trickle-bed.

Page 38: Natural Gas Conversion

34

Figure 2. Idealized diffusion-reaction system in a three-phase trickle-bed reactor, where one reactant is in gas-

phase and the other in liquid phase [Gianetto and Silveston, 1986].

2. Dispersion model equations for multiphase systems

Typically, mass and energy balances for a 1-dimensional dispersion model of a trickle-bed

reactor, with A as the reactant of the gas phase and B as that of the liquid phase, would look

like this if the reaction only takes place at solid, and assuming constant superficial gas and

liquid velocities:

MB for component A in gas phase:

0)()( '

2

2

AL

AvL

AGiGL

AGeAGL C

H

paK

dz

dCu

dz

CdD

where ε and εL are the void fraction of packing and liquid hold-up ([m3 liquid/m

3 reactor]),

respectively. (ε - εL) is equal to the gas hold-up, εG, in [m3 gas/m

3 reactor]. uiG is the interstitial

velocity of gas in [m reactor/s] and hence the expression (ε - εL)*uiG is equal to usG, the

superficial gas velocity in [m3 gas/m

2 reactor, s]. If usG is not constant, it should be included in

the differential. av’ is the gas-liquid interfacial area per unit packed volume in [m

2

interface/m3 reactor]. KL is an overall mass transfer coefficient from gas bulk to liquid bulk in

terms of the liquid concentration gradient, according to:

GLL HkkK

111

where kL and kG are mass transfer coefficients from interface to liquid bulk based on

concentration as driving force, and from gas bulk to interface based on partial pressure as

driving force, respectively, in [m3 liquid/m

2 interface, s] and [kmol/m

2, bar, s].

Page 39: Natural Gas Conversion

35

MB for component A in liquid phase:

0)('''

2

2

s

AsALvAlALA

vLAL

iLLAL

eALL CCakCH

paK

dz

dCu

dz

CdD

MB for component B in liquid phase:

0)(''2

2

s

BsBLvBlBL

iLLBL

eBLL CCakdz

dCu

dz

CdD

where εL*uiL is the same as the superficial liquid velocity, usL, in [m3 liquid/m

2 reactor, s]. kl is

the mass transfer coefficient between liquid and catalyst surface in [m3 liquid/m

2 interface, s],

and av’’ is the liquid-solid interfacial area per unit packed volume in [m

2 interface/m

3 reactor].

MB for component A inside catalyst pellet, concentration gradients accounted for:

0)(1

1

2

2

NR

j

jAjs

As

effA rd

dC

d

dD

MB for component B inside catalyst pellet, concentration gradients accounted for:

0)(1

1

2

2

NR

j

jBjs

Bs

effB rd

dC

d

dD

where Deff is the effective diffusivity coefficient inside the catalyst pellet in [m2/s], ρs is the

catalyst density in [kg cat./m3 particle], NR is the number of reactions, αAj is the stoichiometric

coefficient of component A in the jth

reaction, and rj is the reaction rate for reaction j. The

sum of αAj*rj is negative for a component that is consumed. If mass transfer limitation inside

of the pellet can be neglected, for instance when the particle diameter is very small which

often is the case in slurry reactors, the last MB may be simplified with:

MB for component A at catalyst surface, no concentration gradients:

0)1()(1

''

j

NR

j

Ajs

s

AsALvAl rCCak

MB for component B at catalyst surface, no concentration gradients:

0)1()(1

''

j

NR

j

Bjs

s

BsBLvBl rCCak

Page 40: Natural Gas Conversion

36

The mass balances above are typical for any multiphase system. Simplifications are often

made such as assuming plug-flow of the gas. In slurry reactor models, in which the catalyst

particles are entrained with the liquid, the slurry phase is often modeled as a CSTR, while in

trickle beds the liquid phase often is treated as a plug flow. According to Shah and Sharma

[1987], in general, the number of equations that must be set up to solve a multi-phase system

equals the sum of all component balances for each phase plus the number of overall material

balances.

EB for gas phase:

0)(4)()()( '

2

2

gwgw

t

g

lgvapggiGLagL TTad

UTTah

dz

dTcu

dz

Td

where λag is the axial conductivity in the gas phase in [W/m gas, K], ha the heat transfer

coefficient from bulk gas to bulk liquid in [W/m2 interface, K], Ug is an overall heat transfer

coefficient for heat transfer between gas and tube wall in [W/m2 gas-wall interface, K] and

agw is the gas-wall interfacial area per inner tube wall area.

EB for liquid phase:

0)(4)()( '''

2

2

lwlw

t

ls

slvllgvapLLiLLalL TTad

UTTahTTah

dz

dTcu

dz

Td

where λal is the axial conductivity in the liquid phase in [W/m liquid, K], hl the heat transfer

coefficient between liquid and catalyst surface in [W/m2 interface, K], Ul is an overall heat

transfer coefficient for heat transfer between liquid and tube wall in [W/m2 liquid-wall

interface, K] and alw is the liquid-wall interfacial area per inner tube wall area.

EB in particle, intraparticle temperature gradients accounted for:

0)(1

2

2

j

NR

j

jsse rH

d

dT

d

d

While slurry reactors are often considered as isothermal, trickle-beds are not. Therefore it is

necessary to set up energy balances for the different phases. However, according to Jakobsen

[2003], the catalyst particles can often be considered as isothermal, and hence the last EB

could be simplified with:

EB at particle surface, no intraparticle temperature gradients:

0)()1()(1

''

j

NR

j

js

s

slvl rHTTah

Page 41: Natural Gas Conversion

37

However, according to Gianetto and Silveston [1986], commercial trickle-bed reactors are

normally considered as adiabatic, since the heat losses from the reactor are negligible

compared to the heat generated by the reaction, and heat conduction is neglected.

Furthermore, according to Shah [1979] and Gianetto and Silveston [1986], all phases at any

given axial position in the reactor is often assumed to have the same temperature. With these

assumptions, the EB for a trickle bed could be simplified with the following expression:

EB for adiabatic trickle-bed with temperature equality at any z-point in the reactor:

0)()1())((1

j

NR

j

jspggiGLpLLiLL rHdz

dTcucu

Momentum equation

In the case of trickle beds, a momentum equation has to be set up. The standard Ergun

equation was used by Wang et al. [2003] in the case of a Fischer-Tropsch reactor, in which

the liquid phase was assumed just to be a thin film surrounding the catalyst particles.

Froment and Bischoff [1990] give some alternative expressions for calculation of the pressure

drop in trickle beds, based on pressure drops measured for a certain bed with either liquid, δL,

and gas, δG, flow only. The equation below is one example:

666.0log

416.0log

2

2

G

LGL

where δ2 is the two-phase frictional pressure drop. The measured pressure drop, dPtot/dz, is

related to δ2 according to:

LL

tot gdz

dP 2

g is the acceleration of gravity in [m/s2]. The equations above are only valid if usG = usL, since

they are based on the assumption that the gas and liquid are quasi-homogeneous. See Froment

and Bischoff [1990] for further examples. Froment and Bischoff [1990] also discussed that in

trickle-bed operation, the liquid and gas flow rates are much lower than in for instance packed

absorbers, and hence single-flow pressure drop equations could be used as a first

approximation, with the void fraction reduced to account for the liquid hold-up.

Page 42: Natural Gas Conversion

38

3. Special issues in modeling of trickle-bed reactors

3.1 Introduction

What make trickle-beds a bit more complicated to model than other multi-phase reactors, are

the facts that the catalyst is often not completely wetted, insufficient liquid hold-up,

backmixing etc. [Shah and Sharma, 1987]. Gianetto and Silveston [1986] summarized the

main differences between transport and reaction in porous catalysts in multi-phase reactors

and conventional gas-solid reactors with the following points:

- The catalyst pores in multiphase reactors are often filled with liquid, and the diffusion in

liquids is approximately 4 orders of magnitudes lower than that in gases. Concentration of

solutes in most cases will be lower than the concentrations of the gas species. The low

concentrations and the slow diffusion result in a low catalyst surface utilization, which in turn

means that the overall rates are much slower than if the reaction could be performed in gas-

phase.

- When the pores are filled, the heat conductivity is about one order of magnitude higher than

for gas-filled pores. This means that the catalyst particles can be assumed isothermal at each

axial point in the reactor, provided the wetting of the catalyst is sufficient.

- In a multiphase system, the catalyst particle may be exposed to non-symmetrical surface

conditions. Part of the catalyst surface may be in contact with the flowing liquid, another part

in contact with a stagnant liquid, and the rest of the particle might be unwetted. This means

that the solute concentration will vary significantly over the external catalyst surface. When

the wetting is poor, the exothermic reactions can lead to evaporation of the liquid in some

pores. Those pores are hence dried out and completely filled with gas. In these cases, complex

temperature gradients will be present inside of the catalyst.

A poor wetting is also believed to be the major cause of poor effectiveness. The incomplete

wetting leads to non-symmetrical concentration gradients in the catalyst pores and

complicates the prediction of reaction rates and the analysis of kinetic data from trickle-beds.

In order to be able to evaluate the true rate constant, k, from laboratory experiments, it is

necessary to be able to correlate the observed rate constant, kobs, to the true one.

In Table 1, the advantages and disadvantages with trickle-beds as summarized by Shah [1979]

are given.

Page 43: Natural Gas Conversion

39

Table 1. Advantages and disadvantages with trickle-beds [Shah, 1979].

3.2 Catalyst wetting and relating kapp to k

Assuming first order kinetics of the reaction inside a trickle-bed, the apparent rate constant,

kapp, will for instance depend on the thickness of the liquid film surrounding the catalyst

pellets, which in turn is dependent on the liquid flow rate, and on the diffusion resistance

inside the pellets. The kapp will of course be different from (lower than) the true rate constant,

k, as long as the reaction is limited by mass transfer. If the reaction is not truly first order,

different kapp values will be obtained at different conversions.

Bondi [1971], a reference in Froment and Bischoff [1990], derived the following relation for

trickle-beds:

b

Lapp L

A

kk )/(

11 '

where 0.5 < b < 0.7

where A’ is a constant. The higher the LρL/Ω, the closer kapp/k is to 1 for hydrodesulfurization

of a heavy gasoil, which is understandable since the film thickness of liquid surrounding the

particles decreases with increased liquid flow rate, and the wetting of the pellets increases.

kapp/k would however never reach 1 as long as there are intraparticle concentration gradients.

Page 44: Natural Gas Conversion

40

In case of porous packing, two types of wetting can be defined [Shah, 1979]:

- Internal wetting or pore filling. This is often complete due to capillary action.

- External effective wetting, i.e. the amount of the particles’ outside area effectively

contacted by the liquid. Almost all mass exchange between the internal liquid and the

flowing liquid occurs through this area.

Little is known on how the liquid-solid contact efficiency depends upon the gas and liquid

flow rates. Higher gas and liquid loadings can increase the contact. Also a higher viscosity

and lower surface tension increases the contacting efficiency. Satterfield, a reference in Shah

[1979], defined the contacting effectiveness in terms of the apparent rate constant, kapp,

obtained from laboratory experiments in trickle-beds, and the true rate constant, k, obtained in

a CSTR, and found the correlation depictured in Figure 3.

Figure 3. Contacting effectiveness vs liquid loading, as proposed by Satterfield. GL is the superficial liquid flow

rate per unit cross-section (usL) [Shah, 1979].

According to Koros, another reference in Shah [1979], there is little correlation between flow

uniformity and contacting effectiveness. The contacting efficiency may increase even if the

flow uniformity is poor. The interstitial bed geometry, catalyst shape and size play important

roles in the contacting effectiveness. At high liquid mass velocities, high contacting

effectiveness can be achieved even for reactor : catalyst diameter ratios as low as 1.4 : 1. For

some systems, no correlation between the contacting effectiveness and mass velocity can be

found.

Mears [1974], another reference in Froment and Bischoff [1990], meant that correlating the

kapp to k and liquid flow rate (or liquid dynamic hold-up) was not correct, and instead

suggested that kapp should be proportional to the external wetted area of the catalyst, which in

turn depends on parameters such as μL, σL, σL,c etc. This is understood since in conventional

trickle-beds both components in gas and liquid are reactants that must meet and react at the

catalyst surface. The lower the degree of wetting, the smaller the interfacial contact area

between liquid and solid and hence the slower the mass transfer rates of B to the catalyst

active sites. The transport of the gas phase reactant A will however be faster the lower the

Page 45: Natural Gas Conversion

41

degree of wetting, which might actually increase the global reaction rate if A is the limiting

reactant. In the hydroprocessing of oils, a sufficiently large liquid hold-up is essential, since if

this is too low, the whole capacity of the catalyst will not be used, due to incomplete wetting.

In case of Fischer-Tropsch, where all reactants are in the gas phase, the degree of wetting

should primarily affect the mass transfer rates from gas to catalyst surface, probably giving a

higher overall rate the lower the degree of wetting, provided that the reaction rate is diffusion

controlled. Gianetto and Silveston [1986] concluded that in trickle-beds, essentially all

reaction occurs by mass transfer of reactant from the gas-covered part of the catalyst surface.

3.3 Liquid hold-up and mass transfer

Biswas et al. [1988], a reference within Froment and Bischoff [1990], performed experiments

that led to the conclusion that in order for the catalyst particles to be fully wetted, an efficient

liquid distribution at the top is required, and the column-to-particle diameter should exceed 20

– 25 to avoid liquid bypassing along the wall. This is of course of big importance if the

trickle-bed involves reaction between components in the gas and the liquid, in order to fully

utilize the capacity of the catalyst loaded into the reactor.

In the trickle-bed, the liquid holdup, εL, consists of liquid inside the pores (internal liquid

hold-up) and outside of the pores (external liquid holdup). The external liquid hold-up

consists, in turn, of the dynamic hold-up, εL’, and the static hold-up. The static hold-up was

related to the Eötvös number, δL*g*dp2/σL (σL is the surface tension of the liquid in [N/m]),

and it varies between 0.02 and 0.05 (Charpentier et al. [1968], a reference within Froment and

Bischoff [1990]). Froment and Bischoff [1990] presented the following expression for the

dynamic hold-up. Note that it is only dependent on the liquid flow rate and not on the gas

flow rate:

)(

2

23

'

pv

L

Lp

L

pL

L dagdLd

c

where L is the volumetric liquid flow rate in [m3 liquid/s], Ω the cross sectional reactor area

[m2 reactor], μL the liquid viscosity in [kg/m, s] and av is the external particle surface area per

unit of reactor volume in [m2 particle/m

3 reactor]. c, α, β and γ are constants (see Froment and

Bischoff [1990]). Although not discussed by Froment and Bischoff [1990], this dynamic hold-

up could probably find use in a so-called two-zone model, in which the liquid is looked upon

as a flowing fraction and a stagnant fraction, as described below in Backmixing.

Henry and Gilbert, a reference in Shah [1979], claimed that a certain minimum liquid hold-up

is required for 100 % catalyst utilization, as illustrated in Figure 4.

Page 46: Natural Gas Conversion

42

Figure 4. Effect of fluid dynamics on hold-up [Shah, 1979].

The liquid hold-up was claimed to depend on (FrL/ReL)1/3

, where FrL = usL/ρL2dpg and ReL =

usLdp/μL.

There are some correlations that might be used to estimate interfacial areas in the trickle-bed.

One example is the following expression, originally developed for a countercurrent packed

column, for ρLL/Ω = 1.5 kg/m2, s:

18.0

,135.004.0'

Re05.1

L

cL

L

v

v Wea

a

where We = ρLL2dp/Ω

2σL, and σL,c is the critical surface tension above which the packing

cannot be wetted.

Respecting the mass transfer between gas and liquid in trickle-beds, there are some specific

results for cocurrent operation available in literature. For low liquid and gas flows,

Charpentier [1977] (a reference in Froment and Bischoff [1990]) developed the following

correlation for the liquid side mass transfer:

9

'

104.20011.0

AL

LvL

DEak [s

-1]

where DAL is the diffusivity in liquid in [m2/s], assuming that the liquid viscosity is not too far

from that of water, and EL = (ΔPtot/Δz)*usL in [W/m3]. Reiss, a reference in Shah [1979],

found a correlation for the liquid side mass transfer at higher gas and liquid rates (pulsing or

spray flow):

5.0' 173.0 LvL Eak

Figure 5 illustrates the above discussion.

Page 47: Natural Gas Conversion

43

Figure 5. Correlation between liquid side mass transfer coefficient kLaL values for cocurrent downflow in packed

beds. aL in the figure is the gas-liquid interfacial area, i.e. av’. U0L in the figure is the superficial liquid velocity,

i.e. usL [Shah, 1979].

The gas-side mass transfer, kGav’, is also depending on the liquid and gas flow rates which is

illustrated in Figure 6.

Figure 6. Energy correlation for kG. U0 is the superficial liquid velocity (us). ε is the bed voidage. Ψ is the

packing shape coefficient, values are found in Table 2 below [Shah, 1979].

Page 48: Natural Gas Conversion

44

Table 2. Values for porosity and packing shape coefficients used in Figure 6 [Shah, 1979].

For the liquid-solid mass transfer, jD correlations for single-phase flow can be used as a first

approximation. These correlations are given in Froment and Bischoff [1990].

Dealing with mass transfer limitations inside of the catalyst pellets in a trickle-bed is done in

the same way as for single-flow fixed beds, i.e. with particle equations or effectiveness

factors. However, the trickle-bed case is much more complex, since in the case of liquid

evaporation (in hydrodesulfurization or hydrocracking) the pores will be filled with both

vapour and liquid, and the theory of the effectiveness factor for such a situation still has to be

worked out.

3.4 Heat transfer

The trickle-bed is characterized by poor heat transfer rates, compared to reactor

configurations in which the liquid is continuous. The transition from homogeneous trickling

flow to pulsed flow correspond to an increase of several hundred percent in the radial heat-

transfer rate. There are not many studies about the heat transfer in trickle-beds. One relation

describing the heat transfer in the trickle-mode was derived by Weekman and Myers, a

reference in Shah [1979]:

GG

l

g

LL

ll

e

k

k

kk

kPrRe172.0PrRe00174.0

71.7

where ke is the effective conductivity of the bed, kl and kg the liquid and gas conductivities.

ReG and ReL are the gas and liquid Reynolds numbers based on superficial velocities and tube

diameter. PrG and PrL are the gas and liquid Prandtl numbers. The overall heat-transfer

coefficient was related to the effective conductivity as:

Z

cucurk

rU

pGsGpLsLt

e

t

)(183.0892.2

[cal h-1

cm-2

K-1

]

where rt is the tube radius in [cm] and usL and usG are in [g/h, cm2]. cpG

* is the specific heat of

saturated air in [cal/g, K]. Z is the axial distance from the inlet in [cm]. These relations were

found for an air-water cocurrent downward system with packing of spheres with sizes

between 38 mm and 65 mm. The radial heat transfer coefficients were found to be much

larger than those observed for single phase liquid flow. The effect of the gas on the heat

transfer was mainly that it increased the velocity of the liquid phase. The radial velocity in the

liquid was larger in two-phase flow.

Page 49: Natural Gas Conversion

45

3.5 Effectiveness factor

The use of effectiveness factors instead of solving the particle equations is convenient.

According to Gianetto and Silveston [1986], the effectiveness factor change much more

slowly through the reactor than do the reaction rate or concentration. Thus, the effectiveness

factor only has to be estimated at the inlet, in the middle and at the outlet of the reactor.

Gianetto and Silveston [1986] reported expressions for the Thiele modulus and effectiveness

factor for various kinetic reaction rate expressions.

However, when the wetting is not complete, conventional intraparticle effectiveness factors

are not applicable since the concentration of reactant at the external catalyst surface is not

uniform. According to Gianetto and Silveston [1986], it is necessary to consider mass transfer

from gas to liquid, from liquid to the portion of the particle in contact with flowing liquid, and

direct transport from gas to the part of the particle in contact with the gas. Expression for an

overall effectiveness factor in case of spherical catalyst particles partly in contact with

flowing liquid, and partly in direct contact with gas, was derived by Gianetto and Silveston

[1986] for first order kinetics. The whole idea is that the overall effectiveness factor, ηo, is

divided into two parts – one for the wetted catalyst surface (f) and one for the “dry” catalyst

surface (1-f), where f is the wetting efficiency:

glo ff )1(

where ηl and ηg are effectiveness factors for entire particle surface (i.e. overall effectiveness

factors that include external mass transfer resistance as well as internal) covered by liquid and

gas, respectively. The particle effectiveness factor, η, i.e. the one concerning internal mass

transfer restrictions only, is included in the expressions for ηl and ηg. See Gianetto and

Silveston [1986] for the complete expression. The overall rate, r, of formation/disappearance

of one component can then be expressed as:

*

lokCr

where Cl* is the liquid phase concentration in equilibrium with the bulk gas concentration.

Partial wetting should only affect the overall rate/effectiveness factor if ηg > ηl. In order to

estimate the role of wetting on the overall rate, it is necessary to know the wetting efficiency f.

A correlation useful for scale-up is:

75.0

,2.005.036.1exp1L

cL

ll WeGaf

where Gal = dp3g/υl

2, and υl the kinematic viscosity of the liquid in [m

2/s].

Page 50: Natural Gas Conversion

46

3.6 Backmixing

3.6.1 One-parameter models

The 1-dimensional dispersion models characterize the backmixing with just a single

parameter – the dispersion coefficient. The assumption that all the mixing processes follow a

Fick’s law type of diffusion becomes increasingly dubious with increasing degree of

backmixing. But the dispersion models are still very widely used due to their simplicity, since

they describe the backmixing (or RTD) by simply one parameter. In multiphase reactors, the

degree of backmixing is accounted for in each phase. The degree of backmixing is often much

higher for the slow moving liquid phase compared to the fast moving gas phase, which often

leads to a plug-flow model (zero dispersion/backmixing) for the latter [Shah, 1979]. The

effective axial dispersion coefficient, Dea, is expressed in dimensionless form as the Peclet

number (based on particle diameter):

ea

pi

aD

duPe

where ui is the interstitial velocity in [m reactor/s] of the phase encountered. The Peclet

number, at a Reynolds number of 10 (based on particle diameter), is approximately 0.2 for the

liquid phase in the trickle-bed, compared to approximately 2 for the fluid in a single-phase

operation [Froment and Bischoff, 1990].

Froment and Bischoff [1990] claimed that in laboratory-scale, the minimum reactor length for

absence of axial dispersion in the gas phase can be an order of magnitude higher for a trickle-

bed than for single-phase fixed beds. However, in industrial reactors, the axial dispersion in

gas phase can be neglected.

Elenkov and Kolev [1972], a reference in Froment and Bischoff [1990], developed the

following expression for estimation of Dea in the liquid in a trickle-bed:

33.0

23

278.0

4068.0

Lv

L

Lv

L

vea

sL

a

g

a

L

aD

u

Deviations from plug-flow in the gas-phase of trickle-bed reactors are often not concerned. Of

course the conventional way to find Dea is to make a tracer experiment in the phase of interest,

and then to calculate the E(t), the residence time distribution function. The E(t) describes the

dimensionless outlet tracer concentration (Cout(t)/Cout(t) integrated from t0 - t) as a function

of time, and by setting up mass balances and boundary conditions / initial values for the tracer

(including the dispersion term), the numerical solution (i.e. Cout(t)) expression may be fitted to

the experimentally found E(t) by choosing a suitable Dea.

According to Shah [1979], maldistribution of the flow is common in laboratory-scale trickle-

beds. The maldistribution, such as channeling, dead zones and bypassing etc. result in

complex RTD evaluation. If such phenomena cannot be avoided in the system, micromixing

models (see Fogler [1999] for an explanation) have to be set up to correlate such unusual

shapes of RTD curves. Micromixing models describe how molecules of different ages

encounter one another in the reactor, unlike the macromixing models (i.e. the RTD) which

only describe the distribution of residence times.

Page 51: Natural Gas Conversion

47

RTDs for each phase is needed in order to describe the backmixing. However, as mentioned

above, the gas phase is often considered to be a plug flow. Figures 7-10 below show an ideal

RTD and RTDs for systems with maldistributional flows for typical trickle-bed reactors.

Figure 7. An ideal RTD curve, and a RTD curve with tailing [Shah, 1979].

Figure 8. RTD curve for a reactor with channeling [Shah, 1979].

Figure 9. RTD curve for a reactor with large dead space [Shah, 1979].

Page 52: Natural Gas Conversion

48

Figure 10. RTD curve for a reactor with bypassing [Shah, 1979].

Channeling and bypassing are undesirable phenomena since they reduce the utilization of the

catalyst. Proper design of the liquid distributor and a calming section in large-scale trickle-

beds can essentially eliminate the maldistributions of flows. It is also important to use the

same type of particles (same porosity) in the large-scale as in the laboratory scale since tailing

will occur due to the static liquid hold-up in the pores, shown in Figure 7.

Shah [1979] presented several models for describing the backmixing/macromixing in trickle-

bed reactors. The most common, although with a rather poor capability of predicting the

backmixing, are the axial dispersion models, as described above in 2. Dispersion model

equations for multiphase systems. The dispersion model gives satisfactory correlation to the

measured RTD data as long as radial-flow nonuniformities are minimized by keeping a large

reactor diameter / particle diameter (> 25). Michell and Furzer, a reference in Shah [1979],

claimed that a dispersion model of the liquid phase does not describe the true nature of

trickling flow. In their so called “bypass model”, the liquid phase is modeled as a laminar film

flow over a series of packing elements with imperfect mixing, or bypassing, at each packing

junction. The liquid film is assumed plug-flow, and the static liquid hold-up at the junctions

well-mixed by hydrodynamic effects, such as ripples, that cause axial dispersion in the

reactor. At each junction, a fraction q of the flow enters the perfectly mixed regions, while the

rest of the flow bypasses the mixing. This is illustrated in Figure 10.

Figure 10. Schematic picture of the bypass model [Shah, 1979].

Page 53: Natural Gas Conversion

49

Michell and Furzer derived the following expression for the mean residence time, tmL, of the

liquid:

MFmL tqtt

where η is the bed height divided by packing diameter, tF is the mean residence time of a

single film and tM the mean residence time in a single mixer. Expression for tF and tM are

given below:

f

p

Fu

dt

sL

pstatLL

Mqu

dt

where uf is the mean flow velocity in the film, εstatL the static liquid hold-up and usL the

superficial liquid velocity in [m3 liquid/m

2 reactor, s]. uf is related to the wetted surface area,

aw, according to:

3/13/2

48

4

L

L

Lw

sLf

g

a

uu

The expression for E(t) for a Dirac pulse tracer input, in case of the bypass model equations,

will be:

Mx

x

M

xx

t

t

t

t

xxx

qqtqtE exp

)!1()!1()!(

)1(!)()1()(

1

1

Since aw and usL often can be estimated from experiments, the only unknown parameter that

has to be found, in order to fit the experimental results to the model, is q. Hence, both the

dispersion and bypass models are typical single-parameter models (Dea and q).

3.6.2 Two-parameter model

Shah [1979] presented three two-parameter models and one three-parameter model for

description of the backmixing in multiphase reactors. Below, one two-parameter will be

shown as an example.

This model was referred to as the two-parameter modified mixing-cell model in Shah [1979],

and as the two-zone model in Froment and Bischoff [1990]. According to Froment and

Bischoff [1990], the deviations from plug-flow in the liquid phase of a multi-phase packed-

bed reactor are mainly caused by insufficient contact between the gas and liquid. The reason

for this phenomenon may be preferential paths in the packing or stagnant zones. These effects

can often not be described satisfactorily by effective diffusion models. A more appropriate

Page 54: Natural Gas Conversion

50

description could be made using a two-zone model. Briefly, in such a model, only a fraction

of the liquid flows through the packing, and the rest of the liquid is considered stagnant. The

stagnant liquid phase is assumed to be well-mixed and exchanges mass with the flowing

fraction of the liquid and with the catalyst particles. Note that in the one-parameter bypass

model described above, no mass transfer between the film and stagnant zones was

encountered.

The mass balance for A in the gas-phase is written as shown in 2. Dispersion model equations

for multiphase systems. The mass balance for A in the liquid phases can be written [Froment

and Bischoff, 1990]:

MB for A in flowing fraction of liquid:

0)()( ''''

s

As

f

ALvl

f

ALA

vL

d

AL

f

ALT

f

ALiL

f

L CCakCH

paKCCk

dz

dCu

where uiL’ is the interstitial velocity of the flowing fraction of the liquid, kl is the mass transfer

between the flowing fraction and the solid, kT the mass transfer between the flowing and

stagnant liquid. The superscripts f and d stand for flowing fraction and stagnant fraction,

respectively.

MB for A in stagnant fraction of liquid:

)()( ' s

As

d

ALl

f

AL

d

ALT CCkCCk

where kl’ is the mass transfer coefficient between stagnant liquid and solid.

MB for A at catalyst:

)1()()( ''' sA

s

As

d

ALl

s

As

f

ALvl rCCkCCak

kT and kl’ contain interfacial areas that are not yet well established.

In order to solve the equations, the flowing fraction of liquid, f, must be determined, so that εLf

can be estimated. It is f and kT that comprise the two parameters needed to fit the experimental

tracer data with E(t). The expression for E(t) can be found in Shah [1979]. f can be estimated

by:

m

i

t

tf 1

where ti and tm are the time when the tracer first appears in the effluent, and the average

residence time of liquid in the reactor, respectively. The variance of the RTD is given by:

mT tkf /2 22

Page 55: Natural Gas Conversion

51

ti, tm and σ2 are easily achieved from the RTD curve, and hence both f and kT can be estimated

rather easily from the RTD experiment.

3.7 Conclusion

There is still not enough knowledge in order to make a model that can predict the different

results obtained in laboratory-scale and large-scale trickle-bed reactors, and hence the design

practice for trickle-beds is still in an early stage of development. The high degree of

backmixing that can arise in multiphase systems is often difficult to describe with a

conventional dispersion model, as discussed above. This is due to the complex phenomena

that often take place, such as droplets breaking loose from the liquid phase and getting

mixed/entrained with the gas flow, or in the case of slurry reactors gas bubbles coalesce and

break continuously. Some liquids will also form foam in the column at certain conditions,

which is not easy to predict. In the multiphase systems the interfacial contact areas between

the different phases is important for the mass and heat transfer since the resistance towards

transfer mainly exist in the interfacial “films”. At certain conditions, these films will break up

and the mass and heat transfer rates may therefore change drastically. [Jakobsen, 2006]

Shah [1979] summarized the problems with RTD and scale-up for trickle-beds with the

following points:

- RTD data from laboratory-scale experiments cannot be used for large-scale units.

Often the prevailing flow regimes are different in different scales.

- In trickle-beds, maldistribution of the flow is common, which makes it difficult to

evaluate RTD data. Maldistribution, such as channeling, dead zones and bypassing,

often exist in laboratory-scale units but not in larger scale.

Page 56: Natural Gas Conversion

52

The Fischer-Tropsch process

4.1 Introduction

The Fischer-Tropsch (FT) synthesis, in which synthesis gas is converted to liquid fuels such

as gasoline and diesel is also called GTL (Gas-To-Liquids). Feedstocks for the synthesis gas

can be coal, natural gas, biomass etc. The FT-process was discovered in the early 1920’s and

has been used from time to time at places short of petroleum, such as Germany during WW2,

and in South Africa during the oil embargo. In the last ten years, the FT-process has gained

new world-wide interest due to the forecasts about diminishing petroleum reserves, and due to

the more stringent emission legislations coming up due to the green-house effect. FT-fuels are

much cleaner than petroleum-based fuels since they are essentially free from sulfur and

aromatics. Furthermore, they can be CO2-neutral if made from biomass.

The commercial-scale FT-plants existing in the world today are three coal-based plants in

South Africa (150 000 barrels1 per day (bpd), Sasol), one natural-gas based plant in South

Africa (23 000 bpd, PetroSA), and one natural-gas based plant in Malaysia (15 000 bpd,

Shell). From this year and within the next 5 years there will be several commercial-scale

natural-gas based FT-plants set up in Qatar, which is a country with huge natural-gas

resources. The companies active in this country are Qatar Petroleum/Sasol Chevron, Shell,

ExxonMobile and ConocoPhillips. Statoil recently erected a natural-gas based FT-

demonstration unit in South Africa, which is a slurry reactor with a capacity of 1 000 bpd. In

Germany, Choren/Shell are building a commercial-scale biomass-based FT-plant. Sweden has

currently one pilot-plant for gasification of biomass in Växjö, and one for gasification of

black liquor (a rest product from the pulp industry) in Piteå. Still, the Swedish energy has not

decided what fuels will be made from the syngas produced at this pilot-plants. A feasibility

study made by Ekbom et al. [2005] showed that approximately 25-30 % of the transportation

fuels in Sweden can be replaced by CO2-neutral fuels (FT-fuel, DME or methanol) if all pulp

mills in Sweden replace their old recovery boilers by new black liquor gasifiers. This option

would still cover the steam and energy need of the pulp mills.

In the Fischer-Tropsch (FT) synthesis, synthesis gas (CO + H2) is converted over Fe or Co-

based catalysts into hydrocarbons and water according to reaction (1) [Sie and Krishna, 1999].

CO + 2H2 -CH2- + H2O ΔHR,298K = -165 kJ/mol (1)

Mainly straight-chain saturated hydrocarbons from CH4 up to heavy waxes may be produced.

Some olefins and oxygenates are also produced. The reaction mechanism is a polymerisation

process in which the building block, or monomer, is the CH2-unit. Due to the step-wise

growth mechanism, the hydrocarbon products will always be a range of products of C1+,

described by the ASF (Anderson, Schulz, Flory) -distribution, shown in figure 1. The chain

growth probability, α, is a common parameter used to characterise a FT product distribution.

A higher α means that the product contains higher hydrocarbons, and hence less CH4. It is

known that the higher the pressure, and the lower the temperature, and the lower the inlet

H2/CO-ratio, the higher is α. The chain growth probability is also dependent on the

characteristics of the catalyst (e.g. pellet size, pore size, active site density, promoters etc.). It

is commonly accepted that the highest yield to diesel (the most promising FT-fuel due to a

1 1 barrel equals 159 dm

3.

Page 57: Natural Gas Conversion

53

very high cetane number realised by the un-branched straight hydrocarbon chains) is achieved

by first making wax, which is then hydrocracked into the diesel fraction (C12 – C20). In this

way you will minimise the formation of shorter byproducts, especially CH4. Hence, the

world-wide FT-research is today focused on how to prepare catalysts that give high α, and

how to increase the dispersion of the Co in order to minimize the amount of expensive Co

needed for a certain productivity. A normal α today is around 0.9 for a wax-producing FT-

process.

The molar usage ratio of H2/CO is about 2 over a Co-based catalyst (see reaction (1)). The

reaction is strongly exothermic, which makes a rapid heat removal necessary from the reactor.

An efficient heat removal is a key point in the design of FT reactors, due to the risk of run-

away, and also due to the temperature sensitivity of the product selectivity.

For Fe-based catalysts, the water-gas-shift (WGS) reaction (reaction (2), [Jess, 1999]) takes

place simultaneously with reaction (1) in the reactor, and hence lowering the usage ratio,

which makes it possible to feed the FT-reactor with a molar inlet ratio of H2/CO of less than

2. Such hydrogen poor synthesis gas is obtained from gasified coal or biomass. The WGS

reaction is known for its equilibrium composition which is located “in the middle” between

the reactants and products. Therefore, the WGS equilibrium is often tuned in in chemical

processes.

CO + H2O CO2 + H2 ΔHR,298K = -41 kJ/mol (2)

However, for the case of low-temperature FT (LTFT), which is used when high waxes are

wanted, the WGS reaction is slow and does often not reach equilibrium [Dry, 1996]. Hence,

no commercial FT-process exists today that can handle, directly in the FT-reactor, such a low

H2/CO-ratio as 1.0, which is the ratio in for instance gasified biomass. Instead, external WGS-

units are used in order to correct the H2/CO-ratio in the syngas before entering the FT-reactor.

The FT-reaction is not limited by equilibrium, as is the case in, for instance, the methanol

synthesis. Over a Co-based catalyst, the reaction products (HCs and water) are not believed to

affect the reaction rate, and hence a high conversion per pass is possible (above 90 % possible

in a slurry phase reactor). Over a Fe-based catalyst, water has shown to limit the reaction rate,

and hence the conversion per pass is rather low and water is knocked out prior to

recirculation. Below, two generalised rate expressions for FT over a Fe- and Co-based catalyst

are presented, respectively [Dry, 2002].

For Fe:

For Co:

Co is often preferred to Fe if the syngas is not poor in hydrogen. This is due to its higher

activity, the lack of WGS-activity (which would otherwise lead to a too high H2/CO-ratio

inside of the reactor), and the fact that the reaction rate is not limited by the presence of water.

What limits the conversion per pass over a Co-based catalyst is actually the ability of heat

Page 58: Natural Gas Conversion

54

removal in the reactor. Several different reactor solutions have been developed for FT during

the years. They will be presented in the next section.

4.2 Different reactor types and reaction conditions

In a FT-plant, the production of syngas, if produced from natural gas, accounts for 60 - 70 %

of the total capital and running costs. Economy of scale is very important for the syngas

production, and therefore the downstream FT-reactors must also have a high capacity.

Therefore, the ease of up-scaling is very important when making the selection of type of FT-

reactor. [Dry, 2002]

The FT-synthesis takes place in a three-phase system. The gas phase contains the reactants

and water vapour and gaseous hydrocarbon products. The higher hydrocarbons compose the

liquid phase, and the catalyst is the solid phase. In some cases, often high-temperature FT

(HTFT), all products are vaporised under reaction conditions and hence there are only two

phases present. The amounts of syngas and product molecules that must be transferred

between the phases are quite large in FT compared to the hydrogen molecules transferred in

the hydrotreatment of petroleum oils in today’s refineries. Therefore, in FT synthesis there is

a demand for a high interfacial mass transfer. [Sie and Krishna, 1999]

A nearly isothermal reactor, i.e. a reactor in which the temperature control is good, is possible

to operate at a slightly higher average temperature than a non-isothermal reactor in which

there are risk of hot spots and hence run-away. A slightly higher temperature gives a slightly

higher reaction rate, and therefore a higher productivity [Schulz, 1999]. Hence, a reactor with

a good temperature control can have a higher productivity per gram catalyst, or put in another

way, can host a more active catalyst.

Many of the early FT-reactors (before and during WW2) were only possible for low

capacities (15 bbl/day) due to inefficient heat removal and required very large cooling

surfaces. Low pressure was used (even down to atmospheric) which required very low gas

flows in order to reach high conversions. The lower gas flows gave poor heat transfer.

[Krishna and Sie, 2000]

Other early FT-reactors had external cooling, i.e. the hot outlet gas or liquids were cooled and

then recycled into the reactor. This solution requires very large recycle streams in order to

remove the generated heat in the reactor, which in turn results in large pressure drops and high

energy consumption for the circulation. [Krishna and Sie, 2000]

High heat exchange rates are today accomplished by forcing the syngas at high linear

velocities through long narrow tubes packed with catalyst particles (tubular fixed-bed reactor,

TFBR) to achieve turbulent flow and each tube cooled by boiling water, or by using a slurry

phase reactor (SPR) in which the catalyst is dispersed in the liquid products and in which very

high heat transfer rates are achieved. For HTFT in which α < 0.71, fluidised catalytic bed

reactors are used which have very good heat transfer and temperature equalisation

characteristics. Due to the risk of heavy products depositing on the catalyst pellets, fluidised

beds cannot be used when high waxes are produced. [Dry, 2002; Sie and Krishna, 1999]

There are currently two FT operating modes: [Dry, 2002; Wender, 1996]

Page 59: Natural Gas Conversion

55

1. High-temperature FT (HTFT); 300 – 350 ºC, Fe-based catalysts, production of

gasoline and linear low molecular mass olefines. Significant amounts of oxygenates

are also produced. Diesel may be produced by oligomerization of the olefins.

2. Low-temperature FT (LTFT); 200 – 240 ºC, either Fe- or Co-based catalysts,

production of high amounts of paraffins and linear products and the selectivity to high

molecular mass linear waxes can be very high. The primary diesel cut and the

hydrocracking of the waxes yield exellent diesel fuels. The primary gasoline cut needs

further treatment to obtain a high octane number.

The reactor operating conditions also depend on the reactor type. The pressure is between 1

and 40 atm (= 4 MPa). The SV is between 100 – 1000 h-1

. [Williams, 2002]

4.2.1 HTFT

The Sasol FT plants existing in 1982 all had circulating fluidized beds (CFBs) for their HTFT

process, see Figure 11 (left picture). The CFB reactors are also called the Synthol reactors. In

CFBs there are two phases of fluidized catalyst – dense and lean. The catalyst particles move

down the standpipe in dense phase while they are transported up the reaction-zone in lean

phase. In order to avoid that the feedgas goes up the standpipe, the differential pressure over

the standpipe must always exceed that of the reaction zone. During operation at high

temperature, carbon is deposited on the catalyst particles (the Fe-based) which lowers the bulk

density of the catalyst and hence the differential pressure over the standpipe. Therefore it is

not possible to raise the catalyst loading in the reaction section in order to compensate for the

normal catalyst deactivation with time on stream. [Dry, 2002]

Sasol performed pilot scale experiments in fixed fluidized beds (FFBs, see Figure 11 (right

picture)) in the 1970s. They discovered that the FFB reactor had a higher performance than

their commercial CFB unit. 1989 a commercial FFB unit was brought on-line successfully.

Between 1995 and 1999, the 16 second generation CFB reactors at Secunda were replaced by

eight FFB reactors called Sasol Advanced Synthol (SAS). [Dry, 2002] The fluid bed FT

synthesis at Sasol is mainly used for production of olefins (C2-C7). [Schulz, 1999]

For the fluidised HTFT reactors, the catalyst pellets can be small (< 100 um), and hence

internal diffusion limitations are expected to be absent even though the catalyst pores are

filled with wax. [Dry, 1996]

The advantages of FFBs over CFBs are: [Dry, 2002; Wender, 1996]]

- 40 % lower construction costs. For the same capacity, the FFB reactor is much

smaller.

- Lower pressure-drop.

- Due to the wider reaction section, more cooling coils can be installed, which increases

the capacity of the FFB reactor due to the more isothermal behaviour

- At any moment, all of the charged catalyst participates in the reaction.

- The lowering of bulk density due to fouling is of less significance in the FFB which

implies that less on-line catalyst removal and replacement with fresh catalyst is

required to maintain high conversions, which lowers the overall catalyst consumption.

- The FFB reactor requires less maintenance due to the lower linear velocities. The CFB

reactors are complex and subjected to considerable erosion at the bends in the reactor.

Page 60: Natural Gas Conversion

56

Figure 11. HTFT reactors [Wender, 1996].

4.2.2 LTFT

In top-fed multitubular fixed-bed reactors (TFBR, see Figure 12 (left picture)), the waxes

trickle down and out of the catalyst bed. In slurry phase reactors (SPRs, see Figure 12 (right

picture)) the waxes accumulate inside the reactors and so the net wax produced needs to be

continuously removed from the reactor. [Dry, 2002]

In the FT plant in Sasolburg that was started 1955, five multitubular ARGE reactors were

installed for wax production. They are still in operation. Each reactor consists of 2050 tubes

(5 cm i. d., 12 m long). They operate at 2.7 MPa and 230 C. The production capacity of each

reactor is 21 000 t/year. An additional reactor was installed in 1987 operating at 4.5 MPa.

[Dry, 2002]

In the Shell plant at Bintuli there are four large multitubular reactors with a capacity of 125

000 t/year each. There are ca. 10 000 tubes in each reactor. In this plant, Co-based catalysts

are used. These are much more active than the Fe-based catalysts used in the Sasol plants.

This implies that the tube diameters of the Shell reactors are narrower in order to be able to

handle the higher rate of heat release. [Dry, 2002]

Due to pressure drop constraints in packed beds, the catalyst pellet diameter used in the TFBR

is often larger than 1 mm. When the catalyst pellet diameter is larger than 0.5 mm,

intraparticle diffusion can be rate limiting (H2/CO = 2, 21 bar, 200 – 240 C, pore diameter =

10 – 65 nm) [Sie and Krishna, 1999]. This problem can however be overcome by using egg-

shell impregnation, in which only the outmost layer of the pellets contains the active metal. In

this way, the diffusion distance for the reactants to the active sites is not governed by the size

of the pellets. Since Co is expensive, it is important that all of the impregnated metal is used

in the reaction. [Espinoza et al., 1999]

Page 61: Natural Gas Conversion

57

In the 1950s, slurry phase reactors (SPRs) as alternative to fixed bed reactors were first

encountered. Not until 1993 the first commercial unit based on a SPR for wax production was

brought on-line by Sasol. Its capacity is ca. 100 000 t/year, which equals the total capacity of

the five ARGE reactors, and the catalyst is Fe-based. SPRs require efficient filtration devices

in order to separate the produced waxes from the catalyst slurry. [Dry, 2002]

The SPR consists of a shell in which cooling coils are present for steam generation. The

syngas is fed to the bottom of the reactor and it rises through the slurry. The slurry consists of

liquid products (mainly waxes) with catalyst particles suspended in it. The reactant gases

diffuse through the liquid phase to the catalyst particles. The heavy products form part of the

slurry and the lighter gaseous products and water diffuse through the liquid. The gaseous

products and unreacted reactants pass through the slurry to the freeboard above the slurry bed

and then to the gas outlet. [Espinoza et al., 1999]

In the SPR, the catalyst pellets can be smaller than in the fixed bed, since they are dispersed in

the liquid phase and hence there are essentially no pressure drop constraints. The smaller

catalyst pellets prevent internal diffusion limitations on the overall rate. However, since the

pellets are suspended in the liquid wax, diffusion of the reactants from the gas phase through

the wax to the solid catalyst may be rate limiting. [Dry, 1996]

One advantage with the multitubular fixed-bed reactor is that there are no problems in

separating the wax product from the catalyst. [Espinoza et al., 1999] The advantages of slurry

phase reactors over multitubular reactors are: [Dry, 2002; Wender, 1996; Espinoza et al.,

1999]

- The cost of the reactor is only 25 % of the cost of a multitubular reactor.

- The pressure drop over the reactor is four times lower which results in lower gas

compression costs.

- Lower complexity, easier scale-up.

- A lower catalyst loading is possible, which results in a four times lower catalyst

consumption per tonne of product.

- The slurry bed can operate at higher temperatures and hence results in higher average

conversions due to the more isothermal behaviour. This makes it easier to control the

product selectivities. The SPR can handle catalysts with higher activities than the

multitubular fixed-bed reactor can.

- On-line removal/addition of catalyst allows longer reactor runs due to frequent catalyst

renewal, which is necessary in order to obtain a steady selectivity profile for a single

reactor.

The periodic replacement of the catalyst in the multitubular reactor is cumbersome and

maintenance and labour intensive, and it causes disturbances in the operation of the plant. The

cooling system, with cooling only on the shell-side of the tubes, gives rise to axial and radial

temperature profiles in the tubes. For maximum reaction rates, a maximum average

temperature is desired. This temperature is however limited by the maximum allowable

temperature peak. This maximum temperature is set in order to avoid catalyst sintering and

fouling. The product selectivities are temperature dependent, why it is desirable to be able to

vary the temperature. In tubular fixed bed reactors, the temperature control is however

severely complicated by the need to avoid exceeding the maximum peak temperature.

[Espinoza et al., 1999]

Page 62: Natural Gas Conversion

58

One disadvantage with a fluidized bed (such as in the SPR) is that if a poison, like H2S,

should enter the reactor, the whole amount of catalyst will be poisoned. In the case of a fixed

bed, only the layers of the catalyst closest to the entrance will be poisoned. [Dry, 2002] To

achieve this, the removal of sulphur for slurry bed application must be very effective. This is

one of the critical aspects for successful commercialisation of SPRs operating with a Co

catalyst. [Espinoza et al., 1999]

Figure 12. LTFT reactors [Wender, 1996].

The slurry phase LTFT process is by many authors regarded as the most efficient process for

FT clean diesel production today. This FT technology is also under development by Statoil

(Norway) for use to convert associated gas at offshore oil fields into a hydrocarbon liquid.

[Schulz, 1999] Shell however claims that their TFBR works as well as a SPR [Presentation at

GasFuel05 in Brugge, Belgium]. Fixed-bed reactors are also more commonly used for

laboratory-scale experiments, due to faster and simpler operation for screening of different

catalysts.

4.3 Scope of the work

The scope of this report is to describe how to model a fixed-bed (also called trickle-bed) for

the LTFT-synthesis. The laboratory-scale of these reactor types are used at NTNU and by

Statoil for screening of catalysts. The kinetic information from these reactors can also be used

to model slurry phase reactors. In the kinetic experiments, the pellet size of the catalyst in the

fixed-bed is between 53 and 90 μm, in order to avoid mass transfer limitations on the

measured rate. The tube diameter is 10 mm, and the bed height approximately 50 mm. In

larger-scale fixed-beds, this pellet size would result in high pressure drop.

For comparison, the modelling of a slurry phase FT-reactor will be discussed briefly.

Modelling of reactors is made in order to be able to scale-up a process from laboratory results,

to extrapolate the performance of a reactor at conditions outside the conditions on which the

Page 63: Natural Gas Conversion

59

model is built, to predict hot-spots and run-away, to predict optimum operating conditions

respecting temperature control, recycling ratio etc.

4.4 Modeling of the trickle-bed FT-reactor

As mentioned above, high heat exchange rates in the TFBR is accomplished by forcing the

syngas at high linear velocities through long narrow tubes to achieve turbulent flow. Each

tube is cooled by boiling water. The wish is to obtain an isothermal reactor, and hence the

reaction heat should be removed by radial transport (mainly conduction). However, the heat

conductivity is rather poor, especially at the tube wall, compared to fluidised beds and this

often results in radial temperature gradients, see Figure 13. These gradients affect the

selectivity and may also lead to increased deactivation of the catalyst. The gradients are

highest near the entrance of the reactor since the reaction rates are highest there, and since the

liquid phase is rather small. Further down in the reactor, the reaction rates are slower due to

depletion of reactants and more and more liquid wax is present, which reduces the radial

temperature gradients. [Krishna and Sie, 2000; Sie and Krishna, 1999]

Figure 13. Concentration and temperature profiles in a reactor of the ‘Mitteldruck ’ Fischer–Tropsch process [Sie

and Krishna, 1999].

The radial heat transport is enhanced by larger catalyst pellets and higher gas velocities. There

must be a compromise between the heat transport/pressure drop and the effectiveness, η, of

the catalyst. The larger the catalyst particle, the larger is the Thiele modulus, which can be

thought of as “a reaction rate” / “a diffusion rate”. A large Thiele modulus will result in a low

effectiveness factor, as shown in Figure 14, which means that the usage of the catalyst particle

is low, i.e. only the outer part of the catalyst particle is taking part in the reaction.

Page 64: Natural Gas Conversion

60

Fig. 14. Catalyst effectiveness η as a function of the Thiele modulus Φ for various cobalt and iron catalysts.

Hydrogen/ carbon monoxide molar ratio 2, P = 2.1 MPa, T = 473–513 K [Sie and Krishna, 1999].

Both the effective radial thermal conductivity, λeff, and the wall heat transfer coefficient, αw,

increase with increased Reynolds number (Rep) (see Figure 15). [Sie and Krishna, 1999]

Figure 15. Effective radial thermal conductivity λeff and wall heat transfer coefficient αw as functions of the

particle Reynolds number Rep. dp is the particle diameter and λG is the thermal conductivity of the gas [Sie and

Krishna, 1999].

The problem with the poorest heat exchange rate in the inlet of the reactor, where the need for

heat removal is largest, can partly be solved by adding liquid in order to make sure that the

complete tube operates in the trickle-flow mode. [Sie and Krishna, 1999] Another way to

reduce the gradients at the inlet of the reactor is to lower the conversion per pass (to 20 – 30

%) and recycle the gas outlet. In this way, linear gas velocities are increased. The ARGE

process mentioned above, combines a low conversion per pass and recycling with higher

reaction temperature and pressure. Compared to the classical once-trough TFBR, the ARGE

concept results in an increase in capacity by a factor 25, a reduction in cooling area by a factor

12, and a reduction of the amount of catalyst and steel by a factor of 7. [Krishna and Sie,

2000]

As mentioned above, in FT synthesis there is a demand for a high interfacial mass transfer,

since capillary condensation of the formed waxes makes the pores filled with liquid. [Sie and

Krishna, 1999] Therefore, the modelling of the FT fixed-bed reactor involves description of

the interactions between complex chemical kinetics and transport phenomena.

According to Wang et al. [2003], only a few models for FT fixed-beds are reported in

literature. Atwood and Bennett [1979], a reference in Wang et al. [2003], made a one-

dimensional, heterogeneous plug-flow model. Bub et al. [1980] developed a two-dimensional,

pseudo-homogeneous plug-flow model. Jess et al. [1999] made a pseudo-homogeneous two-

dimensional model. In case of the pseudo-homogeneous models, the intraparticle diffusion

Page 65: Natural Gas Conversion

61

was not considered. In the heterogeneous model by Atwood and Bennett [1979], the

intraparticle gradients were encountered for by means of an effectiveness factor. According to

Wang et al. [2003], the use of an effectiveness factor instead of solving the catalyst

pellet/particle equations, makes a fundamental analysis of intraparticle diffusion-reaction

behaviour impossible.

4.4.1 Wang et al. [2003]

Wang et al. [2003] developed a one-dimensional heterogeneous model, with assumed plug-

flow in the gas-phase. Only two phases were accounted for - gas and solid/liquid. The catalyst

was Fe-based, and hence active for the WGS-reaction. Mass and energy balances were set up

for the two phases. The concentration of the components in the solid phase was expressed as

concentration in liquid, and hence the pores were assumed to be filled with liquid wax. Liquid

surrounding the pellets was not taken into account, and it was assumed that the mass transfer

restriction through the gas film surrounding the pellets was negligible. Hence, gas-liquid

equilibrium was assumed at the interface between the bulk gas and the liquid filled pellet. A

modified Soave-Redlich-Kwong (SRK) EOS was used to correlate the equilibrium between

the components in the gas bulk and the same components solved in the wax in the catalyst

pores. The description of this EOS is given in Wang et al. [1999]. For the heat transfer,

resistance between solid and gas was assumed to be located in a film around the pellets and

was expressed as a boundary condition for the pellet equation.

Wang et al. [2003] chose to include the reaction rate terms in the MB (and EB) for the gas

phase, however integrated over the volume of the pellet and then divided by the pellet volume

in order to obtain volume-average values. This was done due to the simplification that there

exist no mass transfer resistance between liquid and gas phase, an assumption which means

that the equilibrium values are attained immediately. From Bischoff and Froment [1990] we

have learnt that when interfacial gradients are accounted for, these terms would have been

replaced by the flux between the phases, i.e. by kgav(CAg - CS

AS) (and by hfav(Tss – Tg)). Wang

et al. [2003] did not estimate this kg, but instead simply just assumed equilibrium

concentrations in liquid and gas.

For the gas phase mass transfer, convection and flux from pellets was encountered. For the

energy balance in the gas phase, convective heat transport, heat flux from pellets and

conduction through tube wall were included.

For the solid phase mass transfer, diffusion and reaction was encountered. For the energy

balance, conduction inside pellets (dispersion term) and reaction were included.

The one-dimensional model cannot predict any radial gradients, which according to the

discussion above might be of significant magnitude in the fixed-bed.

Page 66: Natural Gas Conversion

62

Below follows the model equations used by Wang et al. [2003] with the same symbols as

used by Froment and Bischoff [1990]:

Bulk gas phase:

MB: drRdz

CudpR

NR

j

jijs

p

is 2

0 13

)1(3)(

(i = 1, nb of total components involved)

EB: )(4)()1(3 2

0 13 gw

t

RNR

j

jjs

p

g

pgs TTd

UdrH

Rdz

dTcu

p

Initial conditions:

z = 0: ci = ci,0 P = Pin Tg = Tin

Liquid/solid phase:

MB:

NR

j

jijs

is

ei rd

dC

d

dD

1

2

2)(

1

(i = 1, nb of key components involved)

EB:

NR

j

jjs

s

e rHd

dT

d

d

1

2

2)()(

1

Boundary conditions:

ξ = 0: 0d

dCis 0d

dTs

ξ = Rp: iV

i

L

ii xy

)( gs

e

fs TTh

d

dT

where yi and xi are the mole fractions of component i in gas and liquid, respectively, and φL

i

and φV

i are the fugacity coefficients in liquid and gas phase, respectively. This boundary

condition means that component i is in its equilibrium concentrations in the wax and the gas.

Wang et al. [2003] chose to describe the pressure drop in the bed with the classical Ergun

equation:

p

sg

pd

uf

dz

dP2

Page 67: Natural Gas Conversion

63

The Peng-Robinson EOS (equation of state) was used to calculate ρg and hence us. The ideal

gas-law can only be applied at low reactor pressures.

The rather complex kinetic rate expressions used in the model by Wang et al. [2003] can be

found in Wang et al. [2001]. Rate expressions were set up for the products, which included

CH4, paraffins and olefins of different carbon numbers, and CO2 (WGS-reaction). All reaction

rate expressions included the chain growth parameter, α, defined above. The chain growth

parameter used by Wang et al. [2001, 2003] was slightly modified compared to the original

αASF as defined by Anderson, Schulz and Flory. The modified α is not constant, but changes

with carbon number, and also accounts for the so-called olefin readsorption factor (βn). If the

catalyst has a high βn, the probability that a desorbed hydrocarbon chain will readsorb on the

catalyst surface and continue to grow is high. This phenomenon gives higher waxes and hence

higher α. The definition of βn can be found in Wang et al. [2001]. This factor can be

calculated knowing the rate constant for the olefin readsorption reaction, k-6, and the partial

pressures of the hydrocarbon products (paraffins and olefins). The βn however approaches 0

rather quickly with increasing carbon number. Practically, ethene, which is the shortest olefin,

has the highest readsorption ability, and it is often an order of magnitude higher than the

readsorption ability of propene.

651

1

2kPkPk

Pk

HCO

CO

ASF

)1(651

1

2 nHCO

CO

nkPkPk

Pk

The parameters k5 and k6 are rate constants for paraffin formation and olefin desorption,

respectively. The description of k1 was not given in Wang et al [2001], but is probably the rate

constant for consumption of CO.

Wang et al. [2001] characterized the wax by NMR and found it to be approximately n-C28H58

in all experimental runs, why this was also put into the model. Dispersion coefficients for H2,

CO and CO2 in this wax, Di, were taken from literature and could be fitted to expressions as a

function of temperature. Dispersion coefficients for the other components were approximated

by means of diffusivity correlations in infinitely dilute solutions:

6.0)/( iCOCOi VVDD

where V is the mole volume. The effective diffusivity of component i was calculated by

correcting the molecular diffusivities for the porosity and tortuosity of the catalyst pellets:

is

ei

DD

Clearly, this model should only work satisfactorily if the composition of the wax is close to n-

C28H58, and therefore should not be so accurate in predicting reactor operation under

conditions far away from those upon which the model is based.

Page 68: Natural Gas Conversion

64

In order to make the system of equations converge, the surface liquid concentrations in

equilibrium with the bulk gas concentrations were set as guess values for the unknown

concentrations inside the catalyst pellets.

Below, some of the results from the simulation are discussed. Figure 16 shows the

concentration profiles of CO, H2, CO2 and H2O inside of the wax-filled catalyst pores (0 is at

center and 1 at pellet surface).

Figure 16. Concentration profiles of key components in catalyst pellet [Wang et al., 2001].

The further the FT-reaction proceeds, the more water is produced, which is seen in the

increasing H2O concentration towards the middle of the particle. The more water, the more

CO is consumed in the WGS reaction, at the same time as it is consumed in the FT-reaction,

which explains the rapid decrease in CO concentration between 1 and 0.8Rp. This means that

the reaction zone is actually only in 0.2Rp, which is in good agreement with a value of 0.24Rp

reported in literature for catalyst pellets of 3 mm in diameter. The rapid decrease in CO

concentration in the outmost particle shell is also due to the far more severe diffusion

limitations on CO compared to H2. However, since the concentrations of CO2 and H2 are

rather high at this point (0.85Rp), the reversed WGS-reaction will be triggered, and hence

provide some CO for the FT-reaction even far inside of the pellet.

Since both the reaction order with reference to PH2 and PCO are positive, lower partial

pressures of these reactants will result in a lower reaction rate inside of the catalyst pellet

compared to at the surface. Furthermore, because of the higher H2/CO ratio inside of the

particle, see Figure 17, the selectivity to waxes will be lower here.

Page 69: Natural Gas Conversion

65

Figure 17. Variation of H2/CO ratio along the pellet dimension [Wang et al., 2001].

An improved diffusion of the CO would increase the reaction rate as well as the selectivity to

waxes.

The authors calculated the effectiveness factor for several different cases. The effectiveness

factor was calculated by dividing the true pellet-volume-averaged consumption rate of CO

with the pellet-volume-averaged consumption rate of CO if the CO concentration would be as

high as at the surface throughout the whole pellet. Figure 18 and 19 show the effectiveness

factor as a function of temperature and pellet diameter, respectively.

Figure 18. Variation of effectiveness factor with operation temperatures at different pressures [Wang et al.,

2001].

Page 70: Natural Gas Conversion

66

Figure 19. Variation of effectiveness factor with pellet size [Wang et al., 2001].

Figure 20 shows the HC distribution as a function of temperature. Figures 21 and 22 show the

selectivity to CH4 and CO2, and C2+ and C5+, respectively, as a function of pellet size.

Figure 20. Product distribution of hydrocarbons at different temperatures [Wang et al., 2001].

Page 71: Natural Gas Conversion

67

Figure 21. Selectivity variations of CO2 and CH4 with pellet radius [Wang et al., 2001].

Figure 22. Selectivity variations of C2+ and C5

+ products with pellet radius [Wang et al., 2001].

The conclusion is that small catalyst particles favour the formation of the desired products, at

the same time as the reaction rate is higher. However, in industrial applications the pellet size

should not be below 2 mm, in order to avoid a large pressure drop. The solution is to use so-

called egg-shell impregnation, in which the active metal, in this case Fe, is situated only in the

outmost shell of the pellet. The size of the inert core will have to be optimised in order to gain

a high selectivity to wax, at the same time keeping the total activity not so much under the

activity of the uniformly impregnated pellet. Figure 23 shows the effect of the inert core

radius on the selectivity to C5+ for a catalyst pellet with a radius of 1.5 mm.

Page 72: Natural Gas Conversion

68

Figure 23. Effect of inert core radius on C5+ selectivity [Wang et al., 2001].

Figure 24 shows that with an inert core radius of up to 80% of Rp, the reaction rate is actually

not lower compared to the uniform pellet. This indicates that the egg-shell impregnation can

improve the selectivity to waxes, avoiding a significant increase of the reactor volume to

obtain the same productivity as compared to uniform catalyst pellets.

Figure 24. Effect of inert core radius on effectiveness factor [Wang et al., 2001].

The model could satisfactorily predict the performance of different-scale demonstration

processes. The main conclusions from the modelling of Wang et al. [2003] were that a large

tube diameter was unfavourable for the total yield of higher HCs, and that the inner diameter

should not exceed 60 mm. This is a direct effect of the increased bed temperature as a larger

tube diameter implies a poorer heat removal (see Figure 25).

Page 73: Natural Gas Conversion

69

Recycling resulted in thermal stability of the reactor, avoiding the hot spot in the beginning of

the reactor, and also enhanced the yield of desired products. With a high recycle ratio the bed

was almost isothermal. The recycle operation means that the superficial gas velocity is

increased, but the space velocity of the fresh feed is unchanged. The lower overall

temperature decreases the CO-conversion, but the total syngas conversion is actually

increased due to the increase in H2/CO usage ratio (i.e. decrease in WGS-reaction rate).

However, from a practical point of view, the power consumption should always be taken into

consideration when choosing the recycle ratio.

An increase in cooling temperature can increase the syngas conversion, but lowers the yield of

higher HCs. Hence, a lower cooling temperature is recommended. An increase in pressure

favours both conversion and yield to desired products.

Fig. 25. Effect of process parameters on the axial temperature profile in the fixed-bed reactor. Base conditions:

dt=32 mm, L=7.0 m, GHSV=500 h−1

, Rcyc=3.0, Tw=Tg,0=523 K, Pin=25 bar. Fresh syngas composition (%): CO:

30.59; H2: 57.75; CO2: 7.0; N2: 4.08; CH4: 0.58 [Wang et al., 2003].

Page 74: Natural Gas Conversion

70

4.4.2 Jess et al. [1999]

Jess et al. [1999] developed a two-dimensional pseudo-homogeneous model for FT-synthesis

over a Fe-catalyst.

In the model by Jess et al. [1999], only mass transfer by convection (only in axial direction)

and reaction was considered in the MB. In the EB, radial conduction, advective heat transport

in axial direction, and reaction was encountered. Note that Jess et al. considered the radial

temperature gradients and radial heat transfer, while neglecting radial concentration gradients.

Below follows the model equations used by Jess et al. [1999] with the same symbols as used

by Froment and Bischoff [1990]:

MB: j

NR

j

ijs

is rdz

Cud

1

)1()(

(i = 1, nb of total components involved)

EB:

NR

j

jjspgser rHdz

dTcu

dr

dT

rdr

Td

12

2

)()1(1

Initial conditions:

z = 0, 0 r Ri: ci = ci,0 P = Pin Tg = Tin

Boundary conditions:

r = 0, all z: 0dr

dT

All z, all r: 0dr

dC

r = Ri, all z: )( ,,

,

iwiR

er

iwTT

dr

dT

r = Ro, all z: )( ,

,

coolow

w

owTT

dr

dT

where Ri is the tube inner radius, and Ro the tube outer radius, αw,i and αw,o [W/m2,K] the inner

and outer wall heat transfer coefficient, respectively. λw is the heat conductivity of the wall

[W/m,K].

Jess et al. [1999] expressed the consumption rate of CO as the sum of three rates, as shown in

Figure 26.

Page 75: Natural Gas Conversion

71

Fig. 26. Kinetic data of the three main reactions during FT-synthesis on the iron-catalyst (ARGE-Lurgi-

Ruhrchemie) [Jess et al., 1999].

The rate expressions above are not truly kinetic. Rather, they are only valid in the case of

mass transfer limitations inside the catalyst pellets. Hence, the rate expressions describe the

rate at which CO is consumed if diffusion of reactants through the catalyst pores is the rate

limiting step, and are, for pellet diameters > 2.5 mm, only valid at temperatures above 170 ºC.

External diffusion limitations could, according to Jess et al. [1999], be neglected up to 400 ºC,

but this is just a hypothetic value since the catalyst would rapidly deactivate at temperatures

above 260 ºC. These rate expressions corresponds to an effectiveness factor of 0.2 at 250 ºC,

and the activation energies in the expressions are approximately half of the true activation

energies, which is typical in the range of pore diffusion limitations [Fogler, 1999]. In their

pseudo-homogeneous model, Bub et al. [1980] also used “non-kinetic” rate data based on

laboratory experiments with rather large particle diameters. By using the same particle

diameter in the laboratory experiments as used in the pilot-scale reactor to be modeled, the

possible intraparticle mass transfer limitation on the overall rate was not necessary to be

considered.

Unlike Wang et al. [2003], the rate expressions of Jess et al. [1999] do not include the product

selectivity. It was not explained how the selectivity was calculated. Since the model is

pseudo-homogeneous, the diffusion of reactants through the waxes inside the catalyst pores is

Page 76: Natural Gas Conversion

72

baked in in the rate expressions, and in this way intraparticle gradients are somehow

accounted for, but this model should not be so accurate in predicting reactor operation under

conditions far away from those upon which the model is based. For instance, it can only be

applied if the catalyst pellets have exactly the same dimensions as the pellets used in the

“kinetic” measurements.

The object was to make a model that could predict the performance with a nitrogen-rich (50

vol-%) syngas. By producing the syngas by partial oxidation of natural gas with air, instead of

with pure oxygen, the need for the capital intense air separation unit is eliminated. Due to the

high inert concentration, this concept can (or must, in order not to build up inerts or to be

forced to bleed off a large stream) work without recycling of the tail gas and hence does not

need a recycle compressor.

The results of the modeling showed that nitrogen plays an important role in removing the heat

released by the FT-reaction. This led to an optimum tube diameter of 70 mm for a stable and

safe operation, compared to 45 mm for the nitrogen-free syngas. The production rate per tube

was about three times higher than in the case with a nitrogen-free syngas, due to the possible

high once-through conversion. This would decrease the investment cost of multitubular fixed-

bed reactors. However, since nitrogen is diluting the system, a higher total pressure will be

needed in order to preserve the partial reactant pressures, which is important for the kinetics.

The critical temperatures were defined as the temperatures not to be exceeded in order to

avoid a temperature runaway. The estimation of these temperatures can be found in Jess et al.

[1999]. Figure 27 shows how the tube diameter influences the critical temperatures in the

nitrogen-free and rich cases.

Fig. 27. Influence of the diameter of the single tubes of a FT-reactor on the critical maximum temperature for

nitrogen-rich and nitrogen-free syngas (volume rate of syngas according to τe = 25 s) [Jess et al., 1999].

Page 77: Natural Gas Conversion

73

4.5 Modeling of the slurry phase reactor

According to Krishna and Sie [2000], the best reactor choice is the slurry phase reactor (or as

they call it “the bubble column slurry reactor”) for large-scale plants with capacities around

40,000 bpd. It was calculated that for such a large total production, 10 TFBR, 17 slurry

reactors operating in the homogeneous regime or four slurry reactors operating in the

heterogeneous regime are needed. The limiting criterion was a maximum reactor weight of

900 tons.

4.5.1 Different flow regimes and estimation of gas hold-up

The homogeneous regime is characterized by a relatively low superficial gas velocity, usG,

and by small bubbles (db = 1 – 7 mm). When the superficial gas velocity reaches the value

Utrans, coalescence of the small bubbles occurs and larger bubbles are produced. The regime

with larger bubbles (db = 20 – 70 mm) and higher usG is called the heterogeneous regime.

(Figure 28 shows a schematic of the two regions.) These larger bubbles have a high rising

speed through the column (1 – 2 m/s) in a plug-flow manner. Smaller bubbles are also present

in the heterogeneous regime, and since they are entrained in the liquid phase, they are often

assumed to have the same backmixing characteristics as the liquid phase. When the catalyst

particles are small (< 50 μm), they are well mixed with the liquid, and the slurry phase can be

regarded as a pseudo-homogeneous phase. This assumption, with no catalyst settling, works

well at larger reactor diameters (> 0.5 m) at high gas velocities (usG > 0.2 m/s).

Fig. 28. Homogeneous and churn–turbulent regimes in a gas–liquid bubble column. The figure on the right

shows that increasing the system pressure delays the regime transition point [Krishna and Sie, 2000].

The volume fraction of catalyst in the liquid, εcat,L (εs in Figure 29) strongly affects the gas

hold up, εG (ε in Figure 29).

Page 78: Natural Gas Conversion

74

Fig. 29. Influence of increased particles concentration on the total gas hold-up in columns of (a) 0.1 m and (b)

0.38 m diameter. Air was used as the gas phase in all experiments. The liquid phase was paraffin oil (density,

ρL=790 kg m−3

; viscosity, μL=0.0029 Pa s; surface tension, σ=0.028 N/m) to which solid particles in varying

concentrations were added. The solid phase used consisted of porous silica particles (skeleton density=2100 kg

m−3

; pore volume=1.05 ml g−1

; particle size distribution, dp: 10%<27 μm; 50%<38 μm; 90%<47 μm). The solids

concentration s, is expressed as the volume fraction of solids in gas free slurry. The pore volume of the particles

(liquid-filled during operation) is counted as a part of the solid phase [Krishna and Sie, 2000].

For modeling purposes it is important to distinguish between the gas hold up in large bubbles

and that in the small bubbles. This can be made by instantaneously switching off the gas flow

to a slurry column at time t = 0. The gas hold-up will quickly decrease due to the escape of

large bubbles. After this quick reduction in gas hold-up, the small bubbles escape and the gas

hold-up is further reduced, as illustrated in Figure 30. From this experiment, the voidage of

gas in the dense phase (see Figure 30), εG,df, the gas hold-up in the large bubbles, εG,b, and the

superficial velocity through the dense phase, us,df, can be estimated. us,df is often taken to be

the same as Utrans.

Fig. 30. Dynamic gas disengagement experiments for air/paraffin oil and air/36 vol.% paraffin oil slurry in the

0.38-m-diameter column. The system properties are as given in the legend to Figure 29 [Krishna and Sie, 2000].

Page 79: Natural Gas Conversion

75

It is clearly seen that the higher the catalyst concentration in the slurry phase, εcat,L, the smaller

is the gas hold-up in the small bubbles. This is due to the increased coalescence of small

particles into bigger in the presence of solid particles. At εcat,L > 30 %, the population of small

bubbles is destroyed. The εG,df for a certain εcat,L is essentially constant at usG > 0.1 m/s for the

heterogeneous regime. Furthermore, the εG,df is essentially independent on reactor diameter

(DT). This is very useful for scale-up purposes, since a rather small reactor can be used to

determine the εG,df under actual reaction conditions.

The εG,b, however, of course increases with increasing superficial velocity through the large

bubble phase (usG-us,df), but seems to be independent on εcat,L in the range 0.16 < εcat,L < 0.36.

Unlike the εG,df, the εG,b (εb in Figure 31) is strongly dependent of the reactor diameter. The

εG,b is reduced at large diameters due to the increase in rise velocity, which, in turn is due to

the reduced wall effects at large diameters. Relationships to find the large bubble rise velocity

for different db/DT ratios is found in Krishna end Sie [2000].

Fig. 31. Influence of column diameter on the hold-up of large bubbles b in (a) paraffin slurries and (b) Tellus

oil [Krishna and Sie, 2000].

These relationships must be modified to be valid at higher pressures, as shown in Figure 32.

The modifications are presented in Krishna and Sie [2000]. Briefly, the expression for

calculation of the rise velocity of large bubbles, ub, is corrected for the higher density of a gas

at higher pressure compared to atmospheric pressure. ub for syngas at 30 bars, which is

common in FT, is for instance only 43% of the ub for air at atmospheric pressure, which is

easily understood since the gas densities are 7 kg/m3 and 1.29 kg/m

3 for the pressurized

syngas and the atmospheric air, respectively.

Page 80: Natural Gas Conversion

76

Fig. 32. The influence of elevated pressure on gas hold-up [Krishna and Sie, 2000].

When the ub is known, the εG,b can be calculated through the steps below. εG,df in a pressurized

system can be estimated by:

Lcat

dfGrefG

G

dfGdfG ,

0,,

48.0

,

0,,,

7.01

where εG,df,0 is εG,df at 0% catalyst concentration, ρG,ref is the density of gas at ambient

conditions, and ρG the density of the gas at pressurized conditions. The rise velocity of the

small bubbles, in atmospheric or pressurized systems, can be calculated by:

Lcat

small

smallsmallu

uu ,

0,

0,

8.01

where usmall,0 is the rise velocity at 0% catalyst concentration. This is not much affected by

pressure. The superficial gas velocity through the dense phase is estimated by:

dfGsmalldfs uu ,,

Since ub is known and defined as:

bG

dfssG

b

uuu

,

, )(

we can now calculate εG,b. The total gas hold-up in a pressurized slurry column can be

calculated according to:

)1( ,,, bGdfGbGG

4.5.2 Mass transfer

As mentioned above, in slurry reactors intraparticle diffusion limitations are often negligible,

due to the small pellets used. Mass transfer from gas to liquid can however be rate limiting.

This is often the case with highly active catalysts in high concentrations in slurry reactors

operating in the heterogeneous regime.

Page 81: Natural Gas Conversion

77

The large bubbles represent the major gas throughput. Conventional calculation of mass

transfer rates predicts a very low mass transfer between the gas and slurry phase for large

bubbles. Experiments have, however, shown that the real mass transfer rates are 5 – 10 times

higher than that expected from conventional calculations based on hold-up and bubble size.

The explanation to this is that during the characteristic time for mass transfer from the gas to

the liquid, the bubbles exchange gas with other bubble size classes. In conclusion, the large

bubbles represent the major gas throughput, but gas-liquid mass transfer is mainly determined

by the interfacial area between the small bubbles and the slurry. In most cases, the mass

transfer between gas and liquid is not limiting the overall rate in FT-synthesis. The volumetric

mass transfer (g/l) coefficient for the large bubbles, (kLa)large, can be calculated according to:

refL

L

bG

elL

D

Dak

,,

arg5.0

)(

where DL is the diffusivity in the liquid phase and DL,ref is a constant of 2 * 10-9

m2/s. The

diffusivities of the reactants CO and H2 at 240 ºC are 17.2 * 10-9

and 45.5 * 10-9

m2/s,

respectively. For the small bubbles, the expression has been found to be the same as for the

large bubbles, but with a constant of 1.0 instead of 0.5:

refL

L

dfG

smallL

D

Dak

,,

0.1)(

Note that the mass transfer coefficients kLa, are not the same as kL in Bischoff and Froment

[1990] in which this is a mass transfer coefficient from g/l interface to liquid bulk. kLa used

by Maretto and Krishna in this case is an overall mass transfer coefficient from gas bulk to

liquid bulk, including the interfacial area.

4.5.3. Mixing

The large bubbles concentrate in the middle of the column and carry the liquid upwards. At

the top of the column, the large bubbles escape from the slurry, and the liquid is recirculated.

The larger the reactor diameter, the bigger is the liquid velocity, VL, upwards in the middle

and downwards at the wall. However, by normalizing the liquid velocities by dividing VL(r)

with the liquid velocity at the central line, VL(0), the radial distributions are similar for all

reactor diameters. Hence, the liquid phase axial dispersion coefficient, Da,L, should be

proportional to VL(0). One useful relationship is:

8/1

32/1)(2.0)0(

L

TLg

UgDV

where υL is the kinematic viscosity of the liquid phase. Krishna and Sie [2000] showed that

the backmixing of the slurry phase is actually the same as for low viscosity liquids, such as

water, at the same superficial gas velocity. Measurements of the Da,L have led to the following

simple relationship:

TLLa DVD )0(31.0,

Page 82: Natural Gas Conversion

78

Estimation of Da,L in a commercial scale bubble slurry column with a reactor diameter of 7 m

and a superficial gas velocity of 0.35 m/s yields Da,L = 10 m2/s. This high value of the

dispersion coefficient suggests that the slurry phase is well mixed, and hence can be modeled

as a CSTR.

4.5.4 Heat transfer

The heat transfer is favored by high gas velocities and high εcat,L (εs in Figure 33), which are

typical for the heterogeneous regime.

Fig. 33. Estimate of the heat transfer coefficients to vertical cooling tubes [Krishna and Sie, 2000].

4.5.5 The model

The model, the results of which were briefly described in Krishna and Sie [2000], can be

found in Maretto and Krishna [1999]. A schematic picture of the model is given in Figure 34,

in which U = usG, and Udf = us,df.

Fig. 34. Generalized two-phase model applied to a bubble column slurry reactor operating in the churn–turbulent

regime [Krishna and Sie, 2000].

Page 83: Natural Gas Conversion

79

The slurry was assumed to have the properties of C16H34. The rate expression for the

consumption of syngas was taken from Yates and Satterfield [1991]:

2)1(

2

2

CO

COH

HCObp

papR

where a and b depend on the catalyst used, and are varying with temperature. Note that

according to the FT-reaction stoichiometry, RH2 = 2RCO. The selectivity was described with

the ASF-distribution and only paraffins (CnH2n+2) were assumed, the molar fraction of which

was expressed as:

1)1( n

ASFASFnx

αASF was set to 0.9.

As mentioned above, in the heterogeneous regime, which was modeled in this article, the

larger bubbles (20 – 70 mm) travel up the column at high velocities (1 – 2 m/s) in a plug-flow

manner, resulting in a well-mixed slurry. The smaller bubbles are hence entrained in the

slurry.

A one-dimensional heterogeneous model was made of a slurry reactor with a diameter of 7 m

and height of 30 m. The temperature was 240 ºC and pressure 30 bar. The large bubbles were

modeled as if being in plug-flow, with a superficial velocity of us-us,df, and the slurry phase

was assumed to be well-mixed. The temperature was assumed being constant in the reactor

and the catalyst particles, why no energy balances are needed. The model assumes that

resistance between liquid and catalyst surface is negligible, and also intraparticle gradients

were neglected. The catalyst pellets had a diameter of 50 μm.

MB large bubbles:

LCO

CO

elGCO

elCOL

elGCOdfssGC

H

Cak

dz

Cuud,

arg,,

arg,

arg,,,)(

))((

LH

H

elGH

elHL

elGHdfssGC

H

Cak

dz

Cuud,

arg,,

arg,

arg,,,

2

2

2

2

2 )())((

MB small bubbles:

ZCH

CakCCu LCO

CO

smallGCO

smallCOLsmallGCO

i

GCOdfs

,

,,

,,,,, )()(

ZCH

CakCCu LH

H

smallGH

smallHLsmallGH

i

GHdfs

,

,,

,,,,, 2

2

2

222)()(

Page 84: Natural Gas Conversion

80

where Ω is the cross sectional area of the reactor in [m2], Z is the dispersion height of the

reactor in [m], and superscript i means inlet of reactor. It was assumed that the concentration

of CO in the inlet gas is equal in large and small bubbles and equals CiCO,G. The same was

assumed for H2. The superficial gas velocity through small bubbles, us,df, was assumed

constant in the whole reactor.

MB liquid phase:

0)()(2,,

,,

,,

arg.,

0

arg,

HCOLLCO

o

sLLCO

CO

smallGCO

smallCOLLCO

CO

elGCOZ

elCOL RZCuCH

CakZdzC

H

Cak

02)()(222

2

2

22

2

2

2 ,,

,,

,,

arg.,

0

arg,

HCOLLH

o

sLLH

H

smallGH

smallHLLH

H

elGHZ

elHL RZCuCH

CakZdzC

H

Cak

where εL is the slurry hold-up, equal to 1-εG, since in the case with a slurry reactor the catalyst

particles are included in the slurry. The equations for the liquid phase are typical equations for

a CSTR, i.e. the concentrations in the slurry are assumed constant in the reactor, and also the

concentrations in the small bubbles, which means that the MB for the slurry is an overall MB

for that phase. Since the concentrations in the large bubbles are not constant, the mass transfer

between gas and liquid must be integrated along the reactor in order to fit in in the overall MB

for the slurry phase. The Henry’s constants were used as estimated solubilities for H2 and CO

in the wax, which was assumed to have the characteristics of C16H34, and equal to 2.478 and

2.964 at a temperature of 240 ºC for CO and H2, respectively.

It was also assumed that the contraction of gas volume due to reaction only affects the large

bubbles. The contraction factor, Ф, for 100 % syngas conversion was estimated to –0.648,

assuming 5 % inert in the syngas feed. Hence, the superficial gas velocity through large

bubbles decreases with conversion according to:

)1)(( ,, Xuuuu dfs

i

sGdfssG

where X is the conversion of CO (or H2). The conversion of CO equals the conversion of H2

in this case since the inlet molar H2/CO ratio is the same as the usage ratio in the reactor, i.e.

2. The reason for why this relationship was set up is that the authors only considered mass

balances and distribution between gas and liquid phase for the reactants (and not the

products). Therefore, there is lack of information considering how many moles of gas that

exist at different z-points in the reactor. The expression above estimates this information.

The solution to the equations above was found by means of an iterative procedure. This kind

of solution is typical when differential equations are coupled with algebraic ones, and in this

case most probably means that the concentration in the well-mixed phases are guessed as start

values, and then integration of the gas-phase along the reactor will give new guessed values of

the concentrations in the well-mixed phases, closer to the true ones. The authors mentioned

that it was possible to describe the concentration of one reactant as a function of the other,

which simplified the solution of the equations.

Simulations were made with usG = 0.12 – 0.4 m/s and εcat,L = 0.20 – 0.35. Since the reactor was

assumed to be isothermal, all reaction heat must be removed. This was modeled by assuming

Page 85: Natural Gas Conversion

81

vertical cooling tubes of 50 mm in diameter, with coolant of 230 ºC. The heat transfer

coefficient from slurry to coolant was estimated by:

)Pr(Re1.0 2

GFrSt

where the different dimensionless groups are explained in Deckwer et al. [1982], a reference

in Maretto and Krishna [1999]. The slurry density was calculated from:

LcatpLcat

Sk

LLSl ,,1

where ρp is the catalyst density in [kg/m3 of particle including voids] and ρSk is the catalyst

skeleton density in [kg/m3 of catalyst solids without voids]. The slurry viscosity was

estimated according to:

)5.41( ,LcatLSl

4.5.6 The results

Results from the simulation are shown in Figures 35 - 37. It can be seen that as the superficial

gas velocity, usG (or U in Figure 35), increases, the conversion decreases, but the productivity

increases. That the productivity increases as the space velocity is increased should be due to

an enhanced mass transfer. Practically, a single pass conversion in a slurry phase FT-reactor

should be around 90 % so that no recycling is necessary. For such a high conversion, it is

necessary to keep the superficial gas flow below 0.3 m/s.

Fig. 35. Fischer–Tropsch reactor simulation results: syngas conversion [Maretto and Krishna, 1999].

Page 86: Natural Gas Conversion

82

Fig. 36. Fischer–Tropsch reactor simulation results: total reactor productivity (column diameter DT=7 m;

dispersion height H=30 m) [Maretto and Krishna, 1999].

Increasing the catalyst concentration in the slurry, εcat,L (or εs in Figure 36), of course

increases both conversion and productivity, but the influence of εcat,L is not only because of

the increasing number of active catalyst sites, but also due to its reducing effect on the total

gas hold-up. So, increasing the εcat,L leads to lower εG and hence higher slurry hold-up, εL,

which in turn means that more catalyst can be loaded into the reactor. The higher the

productivity, the higher the number of cooling tubes needed (Figure 37).

Fig. 37. Fischer–Tropsch reactor simulation results: number of 50-mm-diameter cooling tubes [Maretto and

Krishna, 1999].

A sensitivity analysis was also performed to investigate the effect of the mass transfer

coefficient from gas to liquid, and the effect of the Yates-Satterfield kinetic constant a, on the

overall productivity. The superficial gas velocity, usG, was set to 0.4 m/s and the catalyst

concentration in the slurry to εcat,L = 0.25. A 10-fold increase or 3-fold reduction of kLa as

compared with the base case, had a negligible effect on the productivity. However, doubling

the kinetic constant a resulted in a 60 % increase of the productivity. Hence, it was concluded

that the slurry phase reactor at this base case is kinetically controlled.

Page 87: Natural Gas Conversion

83

5. Conclusion/discussion

It seems common for the few models describing FT in fixed-beds (trickle-beds) not to

consider the liquid as a separate phase. This is probably since the liquid essentially just is

existing as a thin film around the catalyst pellets, and hence the gas phase can be seen as the

continuous phase, and the liquid external hold-up, εL, is practically zero. One could of course

include the liquid phase in the model, as would be the case for other gas/liquid/solid systems

such as slurry reactors or more conventional trickle bed reactors (such as used in the

hydrotreatment of oils) in which the liquid hold-up is much larger and also the liquid flow.

The simplification with no separate liquid phase, in the case of the model by Wang et al.

[2003] who used true intrinsic kinetic rate expressions, is justified if the existence of the

liquid film around the pellets does not affect the overall rate, i.e. if no interfacial gradients

occur between gas and gas/liquid interface, and gas/liquid interface and liquid, and between

liquid and solid. Wang et al. [2003] did account for temperature gradients in a gas film around

the pellets. If mass transfer restrictions in the liquid film around the pellets can be neglected,

there is no need to find the interfacial areas, av´ and av´´, which otherwise seem rather

troublesome to find especially if the catalyst pellets are not completely wetted by the liquid.

The models described above developed by Wang et al. [2003] and Jess et al. [1999] did not

explain how the product yield was calculated. Probably, a material balance over the whole

reactor is needed since no liquid phase is accounted for, and the wax composition is assumed

to be constant, and hence cannot describe the proper product distribution.

In all models presented here, the composition of the wax is assumed to be constant. In order

to develop a model able to predict the performance of a wide range of conditions, the model

should be able to predict the composition of the wax phase. The composition of the wax

determines Dei inside the pores and also the concentrations of components at the gas/liquid

interface at the pore mouths. Maybe, it would be possible to solve such a problem iteratively,

by guessing a composition of the wax, and then integrate the gas phase over the reactor in

order to get a new guess value for the composition. For each composition of the wax, the

equilibrium values for gas-phase concentration and liquid-phase concentration of reactants

and products would be different. In a real FT-trickle-bed, the composition of the wax would

vary along the reactor, and, maybe more important, the thickness of the liquid film will

increase downwards the reactor. Another approach would be to assume that the wax

composition in a z-point is given by the selectivity in that point. In this case, of course the

selectivity is not constant in the reactor, but varies with for instance H2/CO ratio at catalyst

surface, as suggested by Wang et al. [2001].

The fact that the catalyst pores are filled with liquid wax, and that diffusion through wax is

much slower than diffusion through gas, leads to the choice of making a heterogeneous model

to account for interfacial and intraparticle gradients. This could have been made with an

effectiveness factor instead of solving the particle equations. However, such a model would

not be able to predict the product selectivity with a high accuracy outside the conditions used

for building the model, since the selectivity is very much dependent on the H2/CO ratio at the

active site on the internal catalyst surface. Only by solving the particle mass balance the true

H2/CO ratio at the active site can be predicted.

The complexity of the FT synthesis regarding the broad product distribution makes it

desirable to simplify the model equations most often by choosing a 1-dimensional model, and

Page 88: Natural Gas Conversion

84

also to treat the liquid and catalyst as a pseudo-homogeneous phase. It also seems to be

common only to consider the mass balances for the reactants.

Page 89: Natural Gas Conversion

85

References

Aasberg-Petersen, K., Christensen, T., Nielsen, C., Dybkjaer, I., Recent developments in

autothermal reforming and pre-reforming for synthesis gas production in GTL applications,

Fuel. Proc. Tech., 2003, 83, 253-261

Bharadwaj, S., Schmidt, L., Catalytic partial oxidation of natural gas to syngas, Fuel. Proc.

Tech., 1995, 42, 109-127

Biesheuvel, P., Kramer, G., Two-section reactor model for autothermal reforming of methane

to synthesis gas, AIChe Journal, 2003, 49, 1827-1837

Bjørgum, E., researcher at SINTEF, Trondheim, personal communication April 2006

Bub, G., Baerns, M., Bussemeier, B., Frohning, C., Prediction of the performance of catalytic

fixed bed reactors for Fischer-Tropsch synthesis, Chem. Eng. Sci., 1980, 35, 348-355

De Groote, A., Froment, G., Simulation of the catalytic partial oxidation of methane to

synthesis gas, Appl. Catal. A, 1996, 138, 245-264

Dry, M., Practical and theoretical aspects of the catalytic Fischer-Tropsch process, Appl.

Catal. A, 1996, 138, 319-344

Dry, M., The Fischer-Tropsch process: 1950-2000, Catal. Today, 2002, 71, 227-241

Dybkjaer, I., Tubular reforming and autothermal reforming of natural gas – an overview of

available processes, Fuel. Proc. Tech., 1995, 42, 85-107

Ekbom, T., Berglin, N., Lögdberg, S., Black Liquor Gasification with Motor Fuel Production

– BLGMF II – A techno-economic feasibility study of catalytic Fischer-Tropsch synthesis for

synthetic diesel production in comparison with methanol and DME as transport fuels, project

report for contract P13284-1 in the Swedish Energy Agency’s program “Forsknings och

utvecklingsprogrammet för alternativa drivmedel 2003 - 2006 (FALT)”, 2005 (The report is

available at www.nykomb.se)

Espinoza, R., Steynberg, A., Jager, B., Vosloo, A., Low-temperature Fischer-Tropsch

synthesis from a Sasol perspective, Appl. Catal. A, 1999, 186, 13-26

Fogler, H. S., Elements of Chemical Reaction Engineering, 3rd

ed., Prentice-Hall

International, United States of America, 1999

Froment, G., Bischoff, K., Chemical reactor analysis and design, 2nd

ed., John Wiley & Sons,

Inc., United States of America, 1990

Gianetto, A., Silveston, P., Multiphase chemical reactors: theory, design, scale-up,

Hemisphere Publishing Corporation, United States of America, 1986

Page 90: Natural Gas Conversion

86

Godfrey, K., Shackcloth, B., Dynamic modeling of a steam reformer and the implementation

of feedforward/feeback control, Measurement and Control, 1970, 3, T65-T72

Hoang, D., Chan, S., Modeling of a catalytic autothermal methane reformer for fuel cell

applications, Appl. Catal. A, 2004, 268, 207-216

Huang, X., Varma, A., McCready, M., Heat transfer characterization of gas-liquid flows in a

trickle-bed, Chem. Eng. Sci., 2004, 59, 3767-3776

Jakobsen, H., professor at Department of Chemical Engineering, Norwegian University of

Science and Technology (NTNU), personal communication, 2006-03

Jakobsen, H., Fixed Bed Reactors, lecture notes in subject: SIK2053, Reactor Technology,

spring 2003, Department of Chemical Engineering, Norwegian University of Science and

Technology (NTNU), 2003

Jess, A., Popp, R., Hedden, K., Fischer-Tropsch-synthesis with nitrogen-rich syngas –

fundamentals and reactor design aspects, Appl. Catal. A, 1999, 186, 321-342

Kirk-Othmer Encyclopedia of Chemical Technology, Volume 13, 4th

ed., John Wiley & Sons,

USA, 1995

Krishna, R., Sie, S., Design and scale-up of the Fischer-Tropsch bubble column slurry

reactor, Fuel Proc. Tech., 2000, 64, 73-105

Lødeng, R., researcher at SINTEF, Trondheim, personal communication April 2006

Lögdberg, S., Modeling of trickle-bed reactors – The Fischer-Tropsch reaction, examination

assignment in “TKP 4145 Reaktorteknologi” at NTNU, Trondheim, spring 2006

Maretto, C., Krishna, R., Modelling of a bubble column slurry reactor for Fischer-Tropsch

synthesis, Catal. Today, 1999, 52, 279-289

Moulijn, J., Makkee, M., van Diepen, A., Chemical Process Technology, John Wiley & Sons

Ltd, England, 2001

Ochoa Fernandéz, E., Rusten, H., Jakobsen, H., Rønning, M., Holmen, A., Chen, D., Sorption

enhanced hydrogen production by steam methane reforming using Li2ZrO3 as sorbent:

sorption kinetics and reactor simulation, Catal. Today, 2005, 106, 41-46

Rydén, M., PhD student at Chalmers, Gothenburg, Sweden, personal communication March

2006

Schulz, H., Short history and present trends of Fischer-Tropsch synthesis, Appl. Catal. A,

1999, 186, 3-12

Page 91: Natural Gas Conversion

87

Shah, Y., Gas-Liquid-Solid Reactor Design, McGraw-Hill Inc., United States of America,

1979

Shah, Y., Sharma, M., Gas-Liquid-Solid Reactors, Ch. 10 in Chemical reaction and reactor

engineering, Marcel Dekker, Inc., 1987

Sie, S., Krishna, R., Fundamentals and selection of advanced Fischer-Tropsch reactors,

Appl. Catal. A, 1999, 186, 55-70

Stehlik, P., Mathematical model of a steam-reforming radiation chamber - V. Effect of

selected parameters on the main characteristics values of the process, Chemicky Prumysl,

1991, 41, 23-28

Ullman’s Encyclopedia of Industrial Chemistry, Volume A 12, 5th

ed., VCH, Germany 1989

Wang, Y., Li, Y., Bai, L., Zhao, Y., Zhang, B., Correlation for gas-liquid equilibrium

prediction in Fischer-Tropsch synthesis, Fuel, 1999, 78, 911-917

Wang, Y., Xu, Y., Xiang, H., Li, Y., Zhang, B., Modeling of catalyst pellets for Fischer-

Tropsch synthesis, Ind. Eng. Chem. Res., 2001, 40, 4324-4335

Wang, Y., Xu, Y., Li, Y., Zhao, Y., Zhang, B., Heterogeneous modeling for fixed-bed

Fischer-Tropsch synthesis: Reactor model and its applications, Chem. Eng. Sci., 2003, 58,

867-875

Wender, I., Reactions of synthesis gas, Fuel Processing Technology, 1996, 48, 189-297

Wilhelm, D., Simbeck, D., Karp, A., Dickenson, R., Syngas production for gas-to-liquids

applications: technologies, issues and outlook, Fuel Proc. Tech., 2001, 71, 139-148

Williams, Fischer-Tropsch synthesis, Oral presentation by Rachid Oukaci at the 2nd

Annual

Global GTL Summit Executive Briefing, May 28-30, 2002, London, UK.