You are on page 1of 50

CHAPTER 1

INTRODUCTION
Energy demand in the world is increasing at an alarming rate. Energy is the basis of life.
Energy is needed everywhere i.e. production and manufacturing, power communication,
cooling and heating and for many other purposes. Energy accessibility and availably have
become easier because of the infrastructure development. The European "World Energy
Technology and Climate Policy Outlook (WETO)" has predicted that the growth rate of
consumption of primary energy worldwide would be 1.8% per year for the period 2000-
2030 [1]. In year 2013, that growth rate of consumption of global primary energy
worldwide has been reported with an increase by 2.3% [2]. The reserves of fossil fuels
i.e. coal; natural gas, crude oil, etc are being used to meet the increasing demand of
energy but with the emission of green house gases and other pollutants like CO
y
, SO
y
,
NO
y
, C
y
H
y
and ash [3]. Hence, higher consumption of fossil fuels carries environmental,
social and health costs. Therefore, the reserves of fossil fuels are diminishing at great
extent and becoming increasingly expensive. Currently, the level of CO
2
emissions per
capita for developing nations is 20% of that for the major industrial nations. As
industrialization in developing nations increases, CO
2
emissions level will increase
substantially. By 2030, CO
2
emissions from developing nations could rise for more than
half the world CO
2
emissions [2].
Energy security has become a major issue. The reserves of fossil fuel are confined to
a very few areas of the world and continuity of supply of these reserves is governed by
ecological economic and political factors. All these factors are conspiring to force
volatile, often high fuel prices while, simultaneously, environmental policy is demanding
a reduction in toxic emissions and greenhouse gases. A coherent energy strategy is
required, addressing both energy supply and demand, taking account of the whole energy
lifecycle including fuel production, transmission and distribution, and energy conversion,
and the impact on energy equipment manufacturers and the end-users of energy systems.
The problems i.e. high rate of fossil fuels consumption, coupled with environmental issue
and diminishing reverses of fossil fuels are forcing to search a cleaner and efficient
source of energy [4].
Although research on carbon dioxide reforming (dry reforming) was initiated in
1920s, but this subject has become one of the major research areas in recent years as this
route possesses the reduction in CO
2
emissions to the atmosphere, a fact, that has become
nowadays a problem of priority. The endothermicity of this reaction, even higher than
that of steam reforming, could be raised at first as a handicap. However, this
characteristic makes of this process the most interesting among the several reversible
reactions that could be applied for energy storage and its transmission to remote areas.
So, involvement of CO
2
in some large scale energy conversion processes based carbon
dioxide reforming could have significant environmental impact. Reforming of
hydrocarbons and alcohols produces a mixture of H
2
, CO, CO
2
and H
2
O. Hydrogen
produced in this process can be either separated or utilized with CO as syngas for the
production of a number of fuels and chemicals. Reforming reactions occur in vapor
phase. Reforming of alcohols demands additional energy for vaporization as they exist in
liquid form. Hence reforming of hydrocarbons is economical over alcohols.
1.1 PROPANE AS FEED STOCK
Among the various hydrocarbons, propane has become prevalent as a feed stock in
reforming processes due to following reasons:
Propane is a major constituent of Liquefied Petroleum Gas (LPG) [5]. LPG is
inexpensive and easily available in the most of developed areas as infrastructure
for its distribution has been already well built-up. LPG has high hydrogen content
in comparison to other higher hydrocarbon. LPG exhibits important properties as
compared to heavier feed stocks enhancing catalytic resistance to the catalyst
deactivation when carbon gets deposited on the catalyst during the reforming
processes [6].
It is produced in high amount from the natural gas and oil crude refining
processes.
It is easily available as it is a product or byproduct in various petroleum and
petrochemical processes.
It produces hydrogen at much lower temperature than natural gas [7].
Its density is high. For very high pressures of 9 bar, it exists in liquid state.
Therefore, it can be easily stored and distributed and transported for mobile
applications [8,9].
1.2 PRODUCTS
As mentioned above, main product of the reforming process is hydrogen and syngas.
1.2.1 Hydrogen
Hydrogen is a non-toxic and non metallic gas having odor, color and taste at standard
temperature and pressure and consists of two molecules with molecular formula H
2
. It is
a highly flammable a combustible gas and it is used with air for burning purposes at a
very wide range of concentrations from 4 % to 75% by volume. It forms is highly
explosive if its concentration with air and chlorine is 474% and 595% respectively. It
occurs with other elements like oxygen as H
2
O, with carbon as hydrocarbons and its
derivatives and in fossil fuels.
Heat of combustion is approximate 141.8 MJ/kg for hydrogen, while for methane; the
heating value is 55.4 MJ/kg [10] which is lower than heat of the combustion for
hydrogen. The heating values for liquid fuels and coal are lower. Hence, it points that a
plethora of energy can be seen in hydrogen as a fuel.
Hydrogen is not considered as a primary source of energy like natural gas and
coal but as an energy carrier because hydrogen is a clean fuel when it gets split from
other elements but this process requires additional energy. Hence, hydrogen acts as an
intermediate to store and carry energy. The hydrogen produced from nuclear sources and
fossil-based energy conversion systems with capture, and safe storage (sequestration) of
CO
2
emissions, and regenerative hydrogen, are almost completely carbon-free energy
pathways. Therefore, hydrogen is an energy carrier with no impact on environment [11].
Hydrogen has following applications:
Heating value of hydrogen is very high producing water as an end product on
combustion. So it can be used as a fuel having no harmful effect on environment.
Hydrogenation processes demand hydrogen e.g. hydro-treating of petroleum
compounds for removal of sulfur and nitrogen present as an impurity in these
compounds, hydrocracking processes to produce gasoline, petrol diesel etc from
high boiling hydrocarbon mixtures. In food industries, hydrogen is used to make
butter and margarine from hydrogenated edible vegetable oil by removing double
bonds resulting in rise in their melting point and improving resistance to rancid
oxidation.
Hydrogen is used for producing many chemical compounds like ammonia (which
acts as the essential of fertilizer industry), hydrochloric acid, hydrogen peroxide,
methyl alcohol (which is used basis of inks, varnishes and paints) etc.
Hydrogen is demanded as a reducing agent in chemical industries and metal
extraction e.g. to treat mined tungsten to make it pure.
Hydrogen is required in AHW (atomic hydrogen welding). Generally, oxy-acetate
is used for welding but hydrogen is used when high temperature is required.
In steel industries, hydrogen has an application in annealing process. Argon is
mixed with hydrogen to weld stainless steel.
As hydrogen is a light gas, it is installed in meteorologists weather balloons
which are fitted with equipments used to study climate conditions.
Electrical generators use the hydrogen as a rotor coolant.

"Apart from above applications hydrogen gas is also used as a fuel in bulk
quantities for providing energy to space shuttles. Fuel cell technology is fast developing
as a potential source to convert chemical energy stored in hydrogen gas into electricity.
Great amount of research work is going on in this field with application to provide power
in remote areas, hydrogen powered cars etc. Hydrogen is hailed as a fuel of future."








Fig 1: Hydrogen: primary energy sources, energy converters and
applications.






1.2.2 Syngas
Syngas (or producer gas or synthesis gas) is a gaseous mixture primarily of containing
carbon monoxide (CO) and hydrogen with in a wide range of varying amounts of the two.

Uses of Syngas
Syngas often acts a standalone fuel. The syngas energy density is only half of the energy
density of natural gas. Therefore, it is mostly suited for use in producing transportation
fuels and other chemical products. As its unabbreviated name implies that, synthesis gas
is mainly used as an intermediary building block for final production (synthesis) of
various fuels as synthetic natural gas, methanol and synthetic petroleum fuel (dimethyl
ether-synthesized gasoline and diesel fuel). In a purified state, the hydrogen component
of syngas can also be used to directly hydrogen fuel cells for electricity generation and
fuel cell electric vehicle (FCEV) propulsion.

Table 1- Reactions using various CO-H
2
ratio as syn gas.[12]
Synthesis
Process
CO/H
2
Ratio
Catalysts
Temp.
(
0
C)
Pressure
(atm)
Products
Methane
- Ni 250-500 1 Methane
Fischer - Tropch 0.5-2.0 Co, Ni, Fe 180-300 1-30
Paraffinic and hydrocarbons
varying from methane to
waxes, plus small quantities
of oxygenated products
Synol 0.3-0.5 Fe 185-225 15-30
Chiefly straight chain
normal alcohols
Methanol 0.3
ZnO, CuO,
Cr
2
O
3

250-350 100-300
Methanol (CH
3
OH)
Higher alcohol
synthesis
0.5 Fe + alkali 400-500 100-300
Alcohols from C
1
to C
n

Isosynthesis 0.5
ThO
2
, ZnO
Al
2
O
3

400-450 100-300
Saturated branched
hydrocarbons
Oxosynthesis (uses
olefins as additional
starting material)
1.2
Cobalt
[Co(CO)
4
]
2

150-200 150-200

Oxygenated hydrocarbons,
aldehydes and alcohols





Fig. 2 - Various applications of syn gas
1.3 COMMON METHODS OF PRODUCTION OF SYNGAS AND
HYDROGEN
Significant attention has been paid towards the processes which are very efficient in
commercial production of syngas (mixture of H
2
and CO) and hydrogen for meeting the
rising demand of chemical compounds and synthetic fuels. There are various common
routes for production of syngas and hydrogen as mentioned below:
1. From Fossil Fuels:
a.) Steam reforming
b.) Dry reforming
c.) Oxidation reforming
d.) Pyrolysis or gasification of coal

2. From Biomass:
a.) Thermo-chemical process
b.) Biochemical process

1.3.1 Steam Reforming
Steam reforming is a well known and traditionally used commercial technology used for
converting hydrocarbons (mainly lighter hydrocarbons, natural gas, Light naptha and
LPG) into hydrogen and syngas. It is the most widely used process for industrial
reforming in world at present. As its name is indicating, steam is used in this process.
This process is a catalytic reaction process of hydrocarbons and steam favored by high
temperature. In this process, the reforming reaction is usually accompanied by water gas
shift (WGS) reaction and methanation reaction. The three major reactions are as
following:

Steam Reforming: C
3
H
8
+ 3H
2
O 3CO+ 7H
2
H
o
298K
=499 kJ/mol (1.1)
Water Gas Shift: CO+H
2
O H
2
+CO
2
H
o
298K
= -41 kJ/mol (1.2)
Methanation: CO
2
+4H
2
CH
4
+2H
2
O H
o
298K
= -206 kJ/mol (1.3)

The steam reforming is generally achieved in a processing device called reformer
around at a temperature of 950 K over a Ni/Al
2
O
3
catalyst packed into the tubes of
reformer which is basically a reforming furnace. As the reaction is carried out at high
temperature, hydrocarbon cracking also occurs with the above stated reactions. Cracking
reactions can be summed up as:

Propane cracking: C
3
H
8
C
2
H
4
+CH
4
H
o
298K
= 89 kJ/mol (1.4)
Ethylene cracking: C
2
H
4
2C+2H
2
H
o
298K
= -52 kJ/mol (1.5)
Methane cracking: CH
4
C+2H
2
H
o
298K
= 75 kJ/mol (1.6)

As it is clear from the above reactions, carbon formation occurs due to cracking
reactions which gets deposited on the catalyst surface making it deactivated. In water gas
shift reaction, CO is further converted into hydrogen and carbon dioxide. As a result, the
final product stream contains CO
2
and hydrogen as main product. CO
2
is removed by
various methods as discussed later in the text. In some plants after the removal of CO
2
,
methanation reaction (the reverse of methane reforming is used to remove the traces of
carbon oxides.

In the reformer, as the temperature is kept very high, the process gas left the
reformer at high temperature. The heat of end product gas can be utilized as a heat source
for additional reforming. "Reforming catalyst is packed in the tubes of a heat exchanger,
and the primary reformer outlet gas flows in the shell. Here the hot gas from the primary
reformer mixes with the gas leaving the open-ended catalyst tubes and then flows along
the outside of the catalyst tubes."

"The heat exchanger reformer temperature is generally found lower than that of
the primary reformer. The lower temperature can be corrected by increasing steam to
hydrocarbon ratio in the heat exchanger reformer as it affects the reforming equilibrium.
The main purpose of using heat exchanger reformer is to decrease the steam generation
and fuel demand [13].
Research work is being done on improvement of catalyst, heat transfer and
operating conditions but still main disadvantage associated with this process are as
following:
High operating temperature,
Carbon (or coke) formation on surface of catalyst and;
High endothermicity of reaction
Occurrence of water gas shift reaction parallel to reforming reaction giving CO
2

which has a bad impact on environment with respect to global warming.

Apart from above reasons, steam reforming involves an additional cost
represented by steam generation compared to other routes for hydrogen or syngas such as
dry reforming, despite the availability of lower cost water. All of these factors, along with
the original high energy consumption required for such an endothermic reaction, were
anticipated to shift attention towards other routes. On the basis of above stated reasons
the dry reforming process is getting more consideration as an alternate to steam
reforming process.

1.3.2 Dry Reforming
Reforming can also be done by using carbon dioxide. This process is named as 'dry
reforming' as there is no involvement of water/steam and CO
2
is used instead of steam in
reforming reaction given as follows:
C
3
H
8
+ 3CO
2
6CO+ 4H
2
H
o
298K
= 622 kJ/mol (1.7)

The reaction is of industrial interest due to lower H
2
/CO ratio in the end product.
This process consumes abundant ad often negative value CO
2
. However literatures have
reported that reforming by CO
2
is uneconomical or not feasible at industry level.
Therefore, it requires robust catalyst that can sustain the reaction at an industrial scale.
Dry reforming of light hydrocarbons for the production of syngas and hydrogen
has become an area of interest as this reaction has a very important environmental benefit
since both methane and carbon dioxide contribute to greenhouse effect causing global
warming and converting the two into valuable feed stock simultaneously reduce the
atmospheric emission of the two gases. Moreover, in dry reforming process of propane
gives an end product as syngas generally that has composition of H
2
/CO ratio (nearly 1)
acting the most suitable feedstock for production of synthesis fuels (downstream
conversion) by Fischer-Tropsch synthesis unlike in conventional steam reforming, H
2
/CO
ratio obtained in product gas is always greater than 3. However, catalyst deactivation due
to coke deposition on catalyst surface is common for both steam and dry reforming
technologies. Therefore, approaches of development of catalysts which are resistant to
their deactivation due to coke deposition on their surface, using a coking agents, and
combination of oxidation of hydrocarbons with reforming reaction can be adopted to
remove the effects of coke formation during the reforming reactions.

1.3.3 Oxidation Reforming
In this type of reforming, as the name indicates, oxygen is supplied with hydrocarbons to
achieve the reforming reactions, hence called as oxidative reforming process. Oxygen
supplied in the process can be pure form or air is used as oxygen supplier. As, this
process demands the pure form of oxygen, the same is commonly achieved by air
liquefaction process increasing the cost of overall process as liquefaction of air is very
expensive process."However, using pure oxygen gives better selectivity and gets rid of
the wasted volume of inert like nitrogen which make up almost three fourth of
atmosphere. The main drawback of this process is additional equipment and capital cost
along with cooling if liquefaction of air is employed as the method of choice for the
production of pure oxygen.

1.3.3.1 Partial Oxidation Reforming
In partial oxidation (POR), hydrocarbons react with oxygen at high temperature to
produce hydrogen and other compounds namely, CO, CO
2
etc. It is non-catalytic reaction
hence heavier hydrocarbons can be used as a feedstock, since there is no difficulty of
catalyst deactivation with heavy hydrocarbon. In general, for light feeds steam reforming
is economical whereas for heavy feeds, partial oxidation is the only feasible choice.
Partial oxidation of propane over Ni catalysts proceeds via combustion (1.8), followed by
steam (1.1) of dry reforming reactions (1.7).
C
3
H
8
+ 3/2O
2
3CO+ 4H
2
H
o
298K
= -227 kJ/mol (1.8)

Partial oxidation is exothermic in nature which results in lower energy
consumption and this is one of the most important advantages of this process over
conventional steam reforming process. But as other process partial oxidation process has
its limitations stated as below:
the cost of pure oxygen supply for partial oxidation of hydrocarbons,
high temperature in the reactor (>1700K) resulting in formation of hotspots on
reactor material surface, and
possible occurrence of carbon deposition, by the CO reduction reaction,
Boudouard reaction and cracking reactions of hydrocarbons (1.4 to 1.6)

CO reduction: CO+H
2
C+H
2
O H
o
298K
= -131 kJ/mol (1.9)
Boudouard reaction: 2COC+ CO
2
H
o
298K
= -172 kJ/mol (1.10)

The above limitations lead to high cost of reactor material and soot formation.

1.3.3.2 Catalytic Partial Oxidation Reforming (CPOR)
The catalytic partial oxidation uses catalyst for partial oxidation process. Since catalyst is
used, lower temperature will be required for CPOR than POR. Also, light hydrocarbon
should be used as a feedstock for CPOR since it uses a catalyst. Oxygen required is less
than the stoichiometric amount to avoid total oxidation of feedstock. When noble metal
catalyst like Pt, Pd, Ru, and Rh is used hydrocarbon can be converted into hydrogen and
syngas with in milliseconds [14]. Since in CPOR residence time required is low,
conversion rate is high, and the selectivity for hydrocarbon is also high it is better choice
for small scale applications [15].

As oxygen is used in the feedstock in POR, lower steam/carbon ratio is required
to prevent coke formation. Since steam requirement is lower hence production of
hydrogen will also be lower and more CO will be formed. Hence POR is desirable where
more CO is required like synthesis gas for chemical production. Since POR capital cost is
high due to requirement of oxygen production. Hence, it is generally used where oxygen
is already available [13].

1.3.3.3 Autothermal Reforming (ATR)
Autothermal reforming of hydrocarbons is a combination of partial oxidation and
steam reforming processes i.e. exothermic and endothermic reaction. The startup of this
process is very rapid which makes it very attractive.

C
3
H
8
+ H
2
O + O
2
3CO + 5H
2
H
o
298K
= 15 kJ/mol (1.11)

General reaction is given as follows
C
x
H
y
O
z
+ nO
2
+(2x-z-2n)H
2
O xCO
2
+ (2x-z-2n +y/2)H
2
(1.11 (a))

Auto thermal reforming of propane is carried out into two separate reaction zones:
First zone: combustion of a part of propane is achieved in a flame or catalytic
burner which produces a hot product stream having very high temperature.
ATR Second Zone: conversion of propane (unconverted) by steam reforming
process to syngas by the products obtained from the previous combustion zone
there are H
2
O and CO
2
.



Using such a process for hydrogen and syngas production provides a major
advantage that there is no external energy is used as the energy form the exothermic
partial oxidation reaction is provided by endothermic steam reforming reaction. This
results in a lower temperature of auto thermal process. As the temperature is low, there is
production of CO
2
and H
2
as at low temperature water gas shift reaction is favored.

ATR uses group VIII transition metals such as Pt, Rh and Ni as catalysts this
catalyst can also be used for steam reforming and catalytic partial oxidation process.
Since endothermic steam reforming reaction relies on exothermic catalytic partial
oxidation, hence better heat integration is required for higher efficiency. However, this
process is suitable where only hydrogen production is under consideration. The syngas
gas ratio produced in this process is greater than one, estimated at about 1.7 from the
previous equation, which means it is not preferred for the GTL process. This reaction is
ineffective especially given some of the shortcomings of steam reforming.
1.3.4 Pyrolysis of Coal Method
Coal pyrolysis is often used syngas production when used as fuel. In this process,
pyrolysis of impure carbon (coke) is followed by alternating blasts of steam and air
mainly by the following simple paths:
C + H
2
O CO + H
2
H
o
298K
= 323.1 kJ/mol (1.12)
C + O
2
CO
2
H
o
298K
= -394 kJ/mol (1.13)
CO
2
+ C 2CO H
o
298K
= 282.1 kJ/mol (1.14)

The first reaction, between incandescent coke and steam, is strongly endothermic,
producing carbon monoxide (CO) and hydrogen H
2
(syngas). When the coke bed has
cooled to as temperature at which the endothermic reaction can no longer proceed, the
steam is then replaced by a blast for air.
The second and third reactions then take place, producing an exothermic reaction
forming initially carbon dioxide raising the temperature of the coke bed- followed by
the second endothermic reaction, in which the latter is converted to carbon monoxide,
CO. The overall reaction is exothermic, producing mixture of CO and hydrogen. Steam
can be re-injected, then air etc., to give an endless series of cycles until the coke is finally
consumed.
1.3.5 From Biomass
Biomass conversion technologies are generally grouped into thermo-chemical and
biochemical processes.

1.3.5.1 Thermo-chemical Process
Thermo-chemical processes tend to be less expensive because they can be operated at
higher temperatures and therefore obtain higher reaction rates. They involve either
gasification or pyrolysis (heating biomass in the absence of oxygen) to produce hydrogen
or hydrogen rich stream of gas (syngas). They can utilize a broad range of biomass types.
However, it is well known that pyrolysis requires large-scale production which causes
large costs, and this process has not yet been realized on a commercial scale.

1.3.5.1 Bio-chemical Process
In this process, enzyme-based biochemical digesters are used with wet, sugar-based
feedstock. Biomass is converted into hydrogen or syngas by anaerobic phenomenon acted
by enzymes on feedstock. Cellulose feedstock could also be used for the same in the
future with continued improvements in process techniques and systems. At medium
production scale and liquid distribution by tanker truck, current delivered costs of
hydrogen from biomass would be in the $5-7 per kilogram range [16].

1.4 METHODS TO PRODUCE HYDROGEN ONLY
Electrolysis of water
In electrolysis, water molecules are split directly into hydrogen and oxygen
molecules using electricity and an electrolyzer device. The overall electrolysis
reaction is given as:
2H
2
O (l) + e- (from electricity) 2H
2
(g) + O
2
(g)
The two basic types of elecctrolyser used in the same are PEM (use a solid
polymer membrane electrolyte) and alkaline (use a potassium hydroxide
electrolyte).

Photo-electricity of water
Solar energy can be used to split the molecules; being renewable in nature it is
environmentally friendly. As major part of solar spectrum consists of visible light
it is essential to find a process so that visible light can be used to decompose
molecule photo catalytically.

High Temperature Fuel Cells
High temperature fuel cells are operated at sufficiently high temperatures to run
directly on methane. The two common basis used for the same are molten
carbonate (MC) or solid oxide (SO) technologies. The process is termed as
internal reforming. Thus, MC and SO fuel system systems do not need a pure or
relatively pure hydrogen stream as do proton exchange membrane (PEM) and
phosphoric acid (PAFC) systems, but can run directly on natural gas or biogas or
landfill gas. Furthermore, such systems can be designed to produce additional
purified hydrogen as a by-product (e.g. for use as a vehicle fuel), by feeding
additional fuel and then purifying the hydrogen-rich 'anode tail gas' from the fuel
cell into purified hydrogen.

1.5 THERMODYNAMIC ANALYSIS
Thermodynamic analysis of the process at a given condition provides what will be the
equilibrium compositions of different products (desired and undesired). Hence it ensures
that whether an experiment investigation would be worthwhile or not. Therefore, before
running the reformer, it would be advantageous to get the operating conditions which
provide high yield of desired product and suppress the undesired product. Theoretical
conversion is the maximum value one can achieve under the given condition; practically
one can reach the conversion value equal to or less than the theoretical value not more than
that. Knowing the optimum conditions, there is an idea to make suitable catalyst for the
process or to modify the catalyst conditions i.e. stability, selectivity or activity of catalyst
with respect to the process.

The thermodynamic equilibrium calculations can be performed via two approaches:
A.) Stoichiometric and
B.) Non stoichiometric approach.

In stoichiometric approach, a set of stoichiometric reaction is used to describe the
system. In this approach thermodynamic analysis is based on equilibrium constants for
these reactions. Equilibrium constant is written as a function of temperature. Since,
equilibrium constant is the product of equilibrium constant in terms of compositions and P
v

where v is sum of the coefficients of the participating species of the reaction with positive
sign for products and negative sign for reactants. Mathematical expression for equilibrium
constant is given as:
K = K
y
P
v

Now K
y
can be expressed in terms of extent of reaction. And then above equation can
solved for extent of reaction and hence equilibrium composition can be determined.

In non- stoichiometric approach, the biggest advantage is that there is no need to
write all possible chemical reactions which might occur in the system apart from that there
is no need for an accurate estimation of initial equilibrium compositions, easy achievement
of convergence in computation and easy incorporation of adsorption and carrier gas effects.
In this approach the first step is to define a list of chemical compounds using atomic
combination of the elements in the feed (C, H, O) which might coexist in the system at
equilibrium. Using direct minimization of Gibbs free energy the equilibrium compositions
of these products can be derived.
















CHAPTER 2
DRY REFORMING OF PROPANE

Although dry reforming reaction dates back to 1888 [17] and was explored
thoroughly by Fischer and Tropsch in 1928 [18], no commercial application is yet reported
as seen by literatures available. Recently, reaction benefits led to increasing attention under
current global considerations [20].
2.1 IMPORTANCE OF DRY REFORMING OVER OTHER PROCESS
Dry reforming reaction has almost the same energy requirement for steam reforming as
may be seen from the thermodynamically aspect. The most important advantage is the
syngas ratio of 0.7, which is lower than 1, thus suitable for the GTL process [21]. In realty,
the actual ratio for methane steam reforming is reported to be 5 compared to theoretically 3
according to Gaffney, Cline & Assocs., 2001 [22]. This high ratio would be more suitable
for H
2
application rather than the GTL process, while adjusting this ratio for the latter
involves extra cost for H
2
separation from steam [23]. Steam reforming involves an
additional cost represented by steam generation compared to other routes for syngas such as
dry reforming, despite the availability of lower cost water. Moreover, the availability of
CO
2
from the coal energy process for example, and its lower cost compared to steam
generation, supports the increased interest. Edwards et al. (1996) [24] successfully
implemented dry reforming of methane in solar energy storage and transport as may be seen
from Fig. 3
It has been also reported that the cost of dry reforming is 20% lower than any other
reforming process [21, 25]. This is may be due to the presence of CO2 as an oxidizing
agent, as Michorczyk and Ogonowski reported that under similar conditions the dry
reforming temperature is lower than simple propane hydrogenation by 100 K and by 50 K
compared to steam reforming [26]. The CO
2
showed excellent activity as an oxidizing
agent compared to H
2
or even O
2
, as the latter may alter catalyst characteristics at high
temperature [27]. Moreover, some crude natural gas resources may contain up to 30%
CO
2
by volume [28]. Biogas, which is also considered as an attractive renewable energy
source, consists of 50-60% CH
4
, 40-50% CO
2
, and moisture . Hence, it is more
economical to utilize it for syngas production instead of adding costs for separation. Dry
reforming is also supported by other advantages such as its suitability for high purity O
2

production [29] as well as usage in chemical energy storage processes [30]. Last but not
least, the environmental contribution of this reaction by consuming

Fig 3 - Concepts for (a) closed and (b) open loop thermochemical heat-pipes based on
methane dry reforming and solar energy, adapted from [48].
green-house gases (GHG), makes dry reforming more preferable under the current
concern for global warming as CO
2
consumption. In a symposium for strategies for the
utilization of fossil fuels in the 21
st
century, urgent action about CO
2
emissions in addition
to other GHGs was recommended [32].
In view of these advantages, dry reforming is considered as the main reaction for
study to find some results that would help in commercializing this important process.
Commercial application was not viable even with the previous benefits for dry
reforming, because of the severe carbon deposition. However, the potential for the
process is beginning to be considered in the context of the vast developments in catalyst
design and reactor operation technologies. The deposited carbon mainly comes from the
hydrocarbon substrate decomposition; hence, the primary routes for the synthesis gas
production under each route may be subjected to different hydrocarbon substrates. Yet,
the commercial application relies on the cost of reactants along with the energy
consumption and operational costs for these routes. Under dry reforming, precise
comparison for the prices of hydrocarbon substrates is mainly unfeasible. This is due to
fluctuations in the price for natural gas and crude oil, the main source for light alkanes
such as propane, even if it is generally accepted that the cost of propane or liquefied
petroleum gas (LPG) is only a small increment.
2.2 PROPANE DRY REFORMING
Very limited studies in the open literature reported propane dry reforming [33-37]
reflecting the new direction in research by considering higher hydrocarbon substrates, as
most of these studies were conducted in the last decade.
In fact, the propane dry reforming reaction given by Eq. 1.7 may be represented
mainly by two steps: namely propane dehydrogenation to hydrogen, methane and
carbonaceous residue (C
x
H
1-x
):
(1.7(a))
And CO
2

gasification of the carbon deposit via:
(1.7(b))
Resulting in the overall reaction:
(1.7(c))
Since this reaction is limited by chemical equilibrium, lower pressures or higher
temperatures are essential for higher conversion as per Le Chatelier's principle [38].
Under thermodynamic study for this reaction, Wang et al. showed that this reaction was
better to be conducted over a temperature range of 700-1100 K and under atmospheric
pressure [39]. Yet, higher temperature involves a thermal cracking reaction to coke and
sintering particularly for Ni based catalysts. Thus, temperatures higher than 757 K and
lower than 972 K may be considered for this reaction under atmospheric pressure.
2.2.1 Reaction Kinetics
The activation energy as well as reaction order may be estimated as:
-r
DR
= k
o1
e
-E/RT
P
a
m
P
b
n

where -r
DR
= rate of propane dry reforming (mol s
-1
g-cat
-1
)
k
o1
= frequency factor (mol s
-1
g-cat
-1
kPa
-(m+n)
)
E = activation energy (kJ mol
-1
)
R
g
= ideal gas constant
T = temperature (K)
p
A
= partial pressure of propane (kPa)
p
B
= partial pressure of carbon dioxide (kPa)
m = reaction order with respect to propane
n = reaction order with respect to carbon dioxide


Table 2- Reported kinetic result for propane dry reforming.
Sr. No. Catalyst T(K) m n
Activation
Energy, E
a

(kJ mol
-1
)
Author
1.
1wt%Ru /Al
2
O
3

873-923 0 0.3 86 Sutton et al.
(2001)
2.
1wt%Rh /Al
2
O
3
, SiO
2
, TiO
2
and MgO
823-923 0 0.4-0.45 - Solymosi et al.
(2003)
3.
2-5wt%Re /Al
2
O
3

923 0 0.6 84 Solymosi et al.
(2005)
4.
1.9 wt% Ni/MgAl(O)
873 - - 93 Olafsen et al.
(2009)
5.
2wt%Ni / Mg(Al)O
873 0.18 0.36 - Jensen et al.
(2009)



Accordingly, in the study carried out by Sutton et al. (2001), over 1wt%Rh /Al
2
O
3

reported zero order for propane and 0.3 for CO
2
[37]. Zero order for propane and fraction
order for CO
2
(0.4-0.45) was also observed by Solymosi et al. (2003) for the Rh catalyst
over various supports; Al
2
O
3
, SiO
2
, TiO
2
and MgO [35]. In 2005, the same group
investigated the propane dry reforming over 2-5wt%Re /Al
2
O
3
and found similar kinetic
results to their previous study and to Sutton et al [36], where the propane order was also
zero and the CO
2
order equalled 0.6. However, the most recent study performed by
Jensen et al. (2009) over 2wt%Ni /Mg(Al)O, the reaction order for propane and CO
2

were 0.18 and 0.36 respectively [38]. Based on the low reaction order values for
propane, it can be assumed that propane was not involved in the rate determining step.
However, since the reaction order mainly depends on the partial pressure and catalyst
activity, the wide range in reactants partial pressures should be examined to determine
the actual reaction order.
2.2.2 Associated reactions
Propane dry reforming reaction is a complex set of reactions under the presence of side
reactions such as propane dehydrogenation, reverse-Boudouard and reverse-water-gas-
shift reaction (RWGS). These side reactions have various influences on the propane dry
reforming reaction path. Thus, this section will provide a brief highlight about each
reaction.

2.2.2.1 Propane dehydrogenation

Although propane dehydrogenation may occur in a series of parallel reactions, all of
them are slightly endothermic. The most likely reaction to take place is:


C
3
H

8

CH
4
+ 2C + 2H
2
H
o
298K
= 37 kJ mol
-1
(1.15)


As per the thermodynamic investigation (39), this reaction is favorable for at higher
temperature than 179 K. The importance of this reaction comes from the production of
surface carbon that deactivates the catalyst and H
2
in addition to various hydrocarbons.
Hence, the catalyst that has higher resistance to carbon deposition and lower selectivity
towards side products would result in higher production of H
2
. Therefore, the appropriate
catalyst should be selected accordingly.
In the propane dry reforming reaction, surface carbon is most likely to be
oxidized by oxygen from CO
2
or even directly. So, using the catalyst that facilitates CO
2

dissociation should be considered. Also, carbon formation can be kept to a minimum by
using some reaction control features such as co-feeding some gases that improve carbon
oxidation. Indeed, carbon formation rate is expected to increase with temperature and
hence, temperature range should be selected carefully, although the effect of carbon
deposition may be modified due to the presence of other reactions occurring
simultaneously.

2.2.2.2 Boudouard reactions


Another side reaction is the Boudouard reaction which also known as CO
disproportionation, where CO may dissociate to produce CO
2
and surface carbon as:


2CO C + CO
2
H
o
298K
= -172 kJ mol
-1
(1.16)



Thermodynamically, this reaction is likely to take place at a reaction temperature
lower than 972 K. However, the presence of high concentrations for CO
2
in the system
may facilitate keeping the influence of this reaction to a minimum. Also, quicker
removal for CO from the system should assist in reducing the effect of this side reaction.
Thus, running under a fluidized-bed reactor may help in removing the products (CO and
H
2
) faster and hence shift the reaction towards product.


2.2.2.2 Reverse-Water-gas shift

This reaction is at equilibrium and may be expressed as:


CO
2
+ H
2
CO + H
2
O , H
o
298K
= 41 kJ mol
-1
(1.17).

Principally, this reaction is very significant in a system containing a rich amount of CO
2

and H
2
like propane dry reforming. It is slightly endothermic and is likely to take place at
a higher temperature than 1073 K. Likewise, prompt removal of H
2
or CO
2
from the
system would lower the contribution of the RWGS reaction. However, since CO
2
is one
of the reactants and would be in excess particularly at a higher feed ratio, R, shifting
CO
2
towards other reactions, such as CO
2
dissociation can also help in this matter.

2.3 Catalyst deactivation
The severe carbon deposition is the main difficulty that prevented a promising reaction
such as propane dry reforming from commercial application. This deposition led to a
loss in catalyst activity with time-on-stream in a deactivation process. The impact of
deactivation on the life time of the catalyst depends on the type of catalytic process as
illustrated in Fig. 4.

In fact, the impact is limited to the catalyst and often extended to include various
consequences that are very important in industries, such as lower production with poorer
quality, further energy requirements and shutdown for catalyst replacement and
associated cost of production time loss and catalyst regeneration [89]. Although the
challenge of deactivation is not totally defeated, various solutions are implemented to
reduce the impact to a lower level through improving catalyst carbon resilience and/or
using some reactor operational strategies [90]. However, to obtain the optimal solution, it
is essential to recognize the nature of the catalyst deactivation process. Hence, this
section underlines the main types of catalyst deactivations that generally occur in
industrial processes, namely poisoning, sintering (aging) and coking (fouling).
Thereafter, this section demonstrates the deactivation sources under propane dry
reforming.



Fig. 4 Time scale of deactivation of various catalytic processes, adapted from [88].








2.3.1 Type of deactivation
2.3.1.1 Poisoning
Poisoning is the loss of activity due to impurities from the feed depositing on the active
portion leading to a decline in the existing active sites for the main reaction. Although
dry reforming is not yet implemented in industrial application, the main cause for
poisoning over Ni-based catalyst surface is the adsorption of H
2
S, As and HCl under a
similar process such as steam reforming [40]. This kind of poisoning is similar to the
trace contaminates in petroleum feed stocks. Poison adsorption in general is a fast
process and the influence depends on the nature of the poison and its amount in the feed.
While sometimes pre-treatment for the feed may reduce the concentration of the poison
to an acceptable level, the strength of adsorption between the poison molecule and
catalyst surface may require higher temperature to crack the interaction [40,41].
Moreover, the tolerance of the metallic active sites for these poisons should be taken into
consideration under the exploration for an active catalyst. However, since Ni based
catalysts are commonly used for propane dry reforming, Fig. 5 exhibits a great tolerance
for these catalysts against sulfur among some of the transition metals such as Ru, Co, Fe
under the methanation process [41].

2.3.1.2 Sintering
This kind of catalyst deactivation is frequently classified as aging, whereupon
structural modification leads to losing catalytic surface and support area. While catalytic
surface area loss may occur as a result of crystallite growth of the catalytic phase, pore
collapse could cause loss of support area. It may take place at any stage in the life cycle of
a catalyst that requires high temperature treatment [42]. Hence, the thermal treatment
needs to be managed cautiously under these stages which include calcination and
reduction during catalyst preparation, reaction operation conditions and regeneration
(carbon burn-off). In supported metal catalysts, sintering occurs usually through two
models as shown in Fig. 6: (a) the atomic migration and (b) the crystallite migration
models. In the first model, metal atoms are captured by larger crystallites during

Fig. 5 Relative steady-state methanation activity profiles for some of the transition
metals as a function of gas phase H
2
S concentration, from [41].
.


Fig. 6 Two theoretical models for crystalline growth under sintering deactivation, (a)
the atomic migration and (b) the crystallite migration, adapted from [42].


relocation from a crystallite to the surface of the support (or in the gas-phase). In the
second model, crystallite collision and coalescence takes place during the migration of
crystallites over the surface of the support

2.3.1.3 Coking
This is a physical type of deactivation (also known as fouling) and reflects the loss of the
catalyst active surface area due to deposition of species obtained from the fluid phase
causing blockage of sites and/or pores [43]. In industrial applications, coking may arise
from various materials such as ashes or soots in the combustion processes, as phaltenes
in petroleum refineries and carbonaceous deposition in many hydrocracking processes
[42]. As seen from the previous section, carbon lay-down is the main complexity that
prevented dry reforming from commercial operation. The process can be described in
Fig. 7 as follow: carbon chemisorbed or physisorbed leading to encapsulating the
metallic active sites with subsequent plugging for micro- and mesopores of the support .
This consequently creates production loss and pressure build-up in the plugged reactor,
hence a costly shutdown is essentially required to resolve the crisis.



Fig 7 Visual illustration for fouling, crystallite encapsulation and pore plugging of
a supported metal catalyst due to carbon deposition, from [41].

2.4 DRY REFORMING CATALYST
Common reforming catalyst consists of one or more metallic active clusters located over
a support material. The support is usually known as an inert material even though
sometimes, metal species may establish a relation with the support which is often
referred to as metal-support interaction. As observed from literature, selecting the
appropriate support is an important matter to minimize the chance of acquiring sintering
deactivation in addition to other factors such as enhancing metal dispersion and creating
an economical catalyst. Extensive research by Bradford and Vannice [44], indicated that
carbon formation is dependent on numerous parameters, such as the metal, metal
crystallite structure, metal-support interactions and support acidity and basicity.

Table 3 displays the most widely used catalysts for dry reforming. It is noticeable that
most catalysts used in these studies were mainly based on group VIII of transition metals (Co,
Ni, Ru, Rh, Pd, Re, Ir and Pt). Indeed, most of the transition metals have been recognized as
good dry reforming catalysts [45]. Among these, the Ni catalyst was generally preferred
commercially for reforming processes. Alumina (Al
2
O
3
) was commonly used as the support
for nickel catalysts because of its high surface area, low cost and thermal stability. Despite the
type of metal, it is found that the combination of metal and support influences the resultant
catalyst activity

Table 3- Metal along with used support in dry reforming process of hydrocarbons
Sr. No. Metal Support
1 Ni
Al
2
O
3
SiO
2
MgO La
2
O
3
CeO
2
ZrO
2
TiO
2

2 Co
Al
2
O
3
SiO
2
MgO TiO
2

3 Rh Al
2
O
3
SiO
2
La
2
O
3
ZrO
2
TiO
2

4 Pt Al
2
O
3
SiO
2
MgO

ZrO
2
TiO
2

5
Ir
Al
2
O
3
SiO
2
MgO La
2
O
3


ZrO
2
TiO
2

6 Pd Al
2
O
3

7 Ru
Al
2
O
3
SiO
2
La
2
O
3


ZrO
2
ZrO
2
Y
2
O
3

CHAPTER 2
PACKED BED REACTOR
PACKED BED REACTOR
Packed bed reactors are the most common reactors in large-scale chemical product
manufacture. They are primarily used for gas-phase reactions in synthesis, combustion,
and effluent treatment [1].
A packed bed reactor consists of a vessel containing one or several tubes of
packed catalyst particles in a fixed, non-mobile bed (Rase, 1990). Different particle
shapes are in use such as spheres, cylinders, rings, flat disc pellets or crushed material of
a certain sieve fraction. Mean particle diameters range from 2 to 10 mm, the minimum
diameter being limited primarily by pressure drop considerations, and the maximum
diameter by the specific outer surface area for mass and heat transfer. Generally, the
gaseous reactant stream passes through these packed tubes, react with the catalyst, and
the product stream leaves from the opposite side. Packed bed reactors are an economical
choice in large scale production. This is due to the fact that they can operate nearly
continuously due to the long catalyst life; which leads to savings in annual costs and
shutdown costs. The first commercial packed bed reactor was made in 1831 by Peregine
Phillips [1]. Phillips patented a process for the oxidation of sulfur dioxide over a platinum
sponge in order to make sulfur trioxide. The sulfur trioxide was then dissolved in water,
making sulfuric acid. Phillips saw the advantage of using a solid catalyst that could be
used continuously without replacement, separation, or recycling. Packed bed reactors are
the dominant type of reactors for reactions requiring solid catalysts. [1]

With regards to reactor application and construction, it is convenient to
differentiate between packed bed reactors for adiabatic operation and those for non-
adiabatic operation. Since temperature control is one of the most important methods to
influence a chemical reaction, adiabatic reactors are used only when the heat of reaction
is low, or where there is only one major reaction pathway; in these cases no adverse
effects are expected on selectivity or yield due to the adiabatic temperature development.
The characteristic feature of an adiabatic reactor is that the catalyst is present in the form
of a uniform packed bed (Fig. 1a) which is surrounded by an outer, insulating jacket.

Reactions with a large heat of reaction, as well as those that are very temperature-
sensitive, are carried out in reactors in which indirect heat exchange occurs with a heat
transfer medium that is circulated through the packed bed. As in most cases the task of
the heat-transfer cycle is to maintain the temperature in the packed bed within a specific
range, this concept is frequently described as an "isothermal packed bed reactor". The
most common arrangement is the multi-tubular packed bed reactor, in which the catalyst
is arranged in tubes, around which the heat carrier circulates externally (Fig. 1b).

Within a production plant the reactor may justifiably be regarded as the central
item of apparatus. However, compared to the remaining parts of the plant for preparing
the feed and for separating and processing the products, it is often by no means the
largest and most cost- intensive component. In many cases, the achievable conversion in
the reactor is limited for thermodynamic (equilibrium) and kinetic reasons (selectivity). It
is then usual to separate the material discharged from the reactor into products and un-
reacted feed components, which are recycled to the feedstock (see Fig. 2). This recycling
procedure involves costs for:
Product separation
Recycle compression
Repeated heating and cooling of the circulating material to the reaction
temperature and back to the temperature of the separating device
Loss of product resulting from the need to remove part of the recycle as a bleed
stream to limit the accumulation of inert substances or byproducts in the recycle
loop.


Fig. 1 Basic types of catalytic packed bed reactors. (a) Adiabatic packed bed
reactor; (b) multitubular packed bed reactor.










Types of Packed Bed Reactor

Adiabatic Reactors
Adiabatic packed bed reactors constitute the oldest configuration of this reactor type. In
the simplest case, they consist of a cylindrical jacket in which the catalyst is loosely
packed on a screen support and is traversed in axial direction (Fig. 9a). To avoid catalyst
abrasion by partial fluidization, random catalyst packings should always be traversed
from top to bottom. If this is not possible, e.g. because of periodic flow reversal ((see
Section 10.1.4.1.1)), special measures must be taken to stabilize the packing. If packed
beds composed of monolith catalyst sections are used, the flow direction can be chosen
freely.
It can be shown [14] that the requirement for a low pressure loss leads to a packed
bed of large diameter and low height (Fig. 9b). Such an arrangement ("disk concept") is
used particularly when very short residence times, followed by direct quenching of the
reaction, are required. Examples include ammonia oxidation in nitric acid production and
oxidative dehydrogenation on silver catalysts (e.g., synthesis of formaldehyde by
dehydrogenation of methanol). In the first case, the "packed bed" consists of several
layers of platinum wire gauze, and in the second case of a porous silver layer several
centimeters in height. The bed diameters can be up to several meters. Flat-bed reactors
with a residence time in the order of milliseconds have recently achieved considerable
attention for partial oxidations of hydrocarbons to synthesis gas, olefins, or oxygenates.

On account of the difficulties involved with obtaining uniform flow as well as for
structural reasons, the disk concept is limited to relatively small catalyst volumes. The
radial flow concept (Fig. 9c) is used when large amounts of catalyst are involved. The
catalyst packing is accommodated in the space between two concentric screens or
perforated plate cylinders, and is traversed in radial direction, either from the inside to the
outside or from the outside to the inside. This design is particularly suitable for large
catalyst volumes as well as for operation at elevated pressure, since at moderate reactor
diameters the catalyst volume can be varied over a wide range by altering the reactor
length, without affecting the flow-through length of the packing and hence the pressure
drop. A critical feature of packed radial-flow reactors is the shape of the upper bed
closure. A simple horizontal covering is not practicable as a gap is formed due to the
unavoidable settling of the packing through which unreacted gas can pass. The
arrangement shown in Fig. 10 has proved effective as it produces mixed axial and radial
flow through the bed in the upper bed closure. The required geometric shape must be
determined by simulation of the local two-dimensional (axisymmetric) flow through the
packing [24].
Monolith catalysts with straight, parallel channels are particularly well suited for
axial flow adiabatic reactors as they allow for a low pressure drop and provide a high
specific outer surface area. Since monolith catalysts are usually produced with a
rectangular cross-section, the packed bed is constructed by arranging these individual
monoliths in rows in the form of a large box. The uniformity of flow into and into and
through adiabatic packed bed reactors is very important. This is not easy to achieve,
particularly with low pressure-loss monolith reactors, and requires a careful design of the
inflow hood. On account of the low pressure loss, unfavorable flow conditions in the
outflow hood may also affect the flow behavior through the catalyst bed. This is of
particular importance for the design and operation of automotive exhaust catalysts, the
inlet or outlet pipes of which often induce a swirl component in the flow, which leads to a
non-uniform flow through the monolith [14].

Purely adiabatic packed bed reactors are used mainly for reactions with a small
heat of reaction or for catalytic combustion. A typical example is gas purification, in
which small amounts of noxious components are converted. Typical applications in the
chemical industry include the methanation of traces of CO and CO
2
in NH
3
synthesis gas,
as well as the hydrogenation of small amounts of unsaturated compounds in hydrocarbon
streams. The latter case requires accurate monitoring and regulation when hydrogen is in
excess, in order to prevent complete methanation due to an uncontrolled temperature run-
away.

Multi-Stage Reactors

In the majority of packed bed reactors for industrial synthesis reactions, direct or indirect
supply or removal of heat in the catalyst bed is necessary to adapt the temperature profile
over the flow path as far as possible to the requirements of an optimal reaction pathway.
Here, a clear historical trend can be observed. This started with the adiabatic reactor (Fig.
12a) which, on account of the adiabatic temperature change, could only be operated to
give a limited conversion. Higher conversions were achieved at the same mean
temperature level when several adiabatic stages were introduced, with intermediate
heating or cooling after each stage. The simplest form involves injecting hot or cold gas
between the stages (Fig. 12b). For a constant tube diameter, the main disadvantages of
this temperature control strategy are cross-sectional loading, which increases from stage
to stage, and the mixing of hot and cold streams, which is energetically unfavorable. The
composition is changed by injection, which can have a positive or negative effect on the
desired reaction. The next step in development was the replacement of injection cooling
by interstage heat exchangers, through which the required or released heat of reaction is
supplied or removed (Fig. 12c). Today, adiabatic multi-stage packed bed reactors with
intermediate cooling or heating are used particularly when the reaction proceeds
selectively to a single product but is limited by the equilibrium conditions. Intermediate
cooling or heating is used to change the gas temperature towards higher equilibrium
conversion. Typical examples include the synthesis of ammonia, sulfur trioxide, and
methanol.

A multi-stage arrangement may be considered for structural, kinetic or economic reasons
in the following cases:
In the case of large single-train plants, subdivision into several individual units is
necessary for reasons of transport or construction.
If a catalyst must be replaced in individual stages at different times on account of
different catalyst deactivation.
If a step-wise addition of reactant has kinetic advantages compared to the total
addition to the feed (here, a suitably designed intermediate heat exchanger ensures
a uniform distribution and mixing within the reaction gas stream).
If the intermediate stages are used to extract a limiting product in the case of
equilibrium-limited reactions; an example is the intermediate absorption of SO
3

before the last stage of the SO
3
synthesis.
With reaction temperatures above 300 C, intermediate cooling can still be
performed directly with boiling water, whereas in a packed bed a high-
temperature heat transfer medium must be used as coolant



Cooled or Heated Multi-tubular Packed Bed Reactors
The development of reactors in which the heat-exchange surfaces are integrated in the
packed bed occurred in parallel with the development of multi-stage adiabatic reactors
with intermediate heating or cooling. The main aim is to supply or remove the heat of
reaction as close as possible to the reaction site. The catalyst is packed into the individual
tubes of the tube bundle. The heat-transfer medium is circulated around the tube bundle
and through an external heat exchanger, in which the heat of reaction is supplied or
removed (Fig. 15). Whereas with endothermic reactions, circulating gas is also in use as
heat-transfer medium, for strongly exothermic reactions exclusively liquid or boiling
heat-transfer media are used. Only in this way can the catalyst temperature be held in the
narrow temperature range necessary for selective reaction control (e.g., in the case of
partial oxidations).

Initially, the integration of heat exchange in the packed bed was utilized to ensure
a reaction being conducted as isothermal as possible. Therefore, reactors of the type
shown in Fig. 15a and b are also commonly referred to as isothermal reactors. They
are comprised of reaction tubes of 20 to 80 mm internal diameter with 0.5 to 15 m length,
and a carefully designed flow control of the liquid heat-transfer medium, with largely
constant heat-transfer conditions and maximum temperature changes of the heat-transfer
medium throughout the tube bundle of a few K. More recent concepts are aimed at
establishing a freely selectable optimum temperature profile over the tube length. This
requires complex heat-transfer medium control with several sections and temperature
levels (Fig. 15c).

Strongly exothermic synthesis reactions such as partial oxidations can only be carried out
in packed bed reactors if the catalyst temperature is controlled in a narrow optimal
window. This can be achieved if the following requirements are fulfilled:
Large heat-transfer areas must be available per catalyst volume (but a small tube
diameter requires small pellet diameters which increase pressure drop).
The temperature of the heat-transfer medium must be close to the desired catalyst
temperature to ensure an effective catalyst temperature control.
A sufficiently large mass flow velocity of the reaction gases is generally
necessary to ensure good heat transport from the packing to the tube walls.

Multifunctional Packed Bed Reactors
Heat-Integrated Reactors
In heat-integrated reactors the hot effluent of a packed bed reactor is used to heat up the
cold feed to the ignition temperature of the catalytic reaction. If the reaction system is all
together moderately exothermic, an auto-thermal operation results where no other
addition or removal of heat is necessary. Autothermal operation always requires a special
start-up procedure to raise the catalyst temperature above the ignition temperature of the
respective reactions. It indicates that an auto-thermal reactor always operates in a region
of multiple steady states in the ignited state.

Heat-Integrated Reactors for Exothermic Reactions
The conventional auto-thermal reactor design consists of an adiabatic packed-bed reactor
coupled with a countercurrent heat exchanger. As mentioned above, the reaction must be
started with help from a separate pre-heater through which the catalyst bed temperature is
raised above the ignition temperature of the reaction. During operation, control measures
must be taken to prevent the reaction from extinction if, for example, the feed to the
reactor is too lean. Because the hot effluent is used to heat up the cold feed, the maximum
temperatures achieved under auto-thermal operation may exceed many-fold the adiabatic
temperature rise.

A sketch of the basic concept, showing an adiabatic packed bed reactor with
feed/exit tubes and two valves for periodic flow reversal, is illustrated in Fig. 27a. Before
starting the operation, the packed bed must again be heated above the ignition
temperature of the catalytic reaction. If the reactor is then fed with cold feed from one
side, the cold feed gas is heated up by the hot catalyst on one side and cooled down by
the cold catalyst on the other side. Two temperature fronts develop and move through the
packed bed. In the first front, the gas is heated up and reacts, but after a certain time
period the direction of flow is reversed via the valves and the temperature fronts are
pushed back.

Heat-Integrated Reactors for Endothermic Reactions
Interesting aspects of autothermal reactor operation arise if an endothermic synthesis
reaction is closely coupled with an exothermic auxiliary reaction in such a way that the
combination is weakly exothermic. If the endothermic reaction has to be run separated
from the exothermic one, a close thermal coupling between both reactions should be
achieved, where the heat of the exothermic reaction must be taken up by the endothermic
reaction as soon as it is set free. In addition, all hot product gases should be used to heat
up all cold feeds.









CHAPTER 4
LITERATURE REVIEW

4.1 EXPERIMENTAL STUDIES
Rama et al. (2014) examined the effect of oxidation pretreatment
temperature (5001000C) on the catalytic activity of Kovar applied on
hydrocarbon CO
2
reforming In this study, catalytic performance evaluation has
been done by using tetra-decane at 800C with 70 mol/s CO
2
that has been
revealed 700 and 1000C as the best pre-oxidation temperature in producing CO
and H
2
, respectively. Changes in composition and morphology brought about by
oxidation pretreatment were determined using X-ray diffraction (XRD) and
scanning-electron microscope (SEM), respectively. XRD and SEM-EDX analyses
showed that a separate metal oxide layer composed of iron oxide (Fe
2
O
3
and
F
3
O
4
), nickel, cobalt, and possibly their respective oxides started to form when
oxidation was conducted at 700C or higher. It has been also observed that the
presence of iron enhanced the stability of nickel in the structure while the
compact structure of Fe
3
O
4
resulted into the formation of a thick and rigid metal
oxide layer on the surface of the Kovar tube. The strong physical bond between
the metal oxide layer and Kovar tube provided the catalyst good mechanical
strength and consequently good catalytic activity.
Siahvashi and Adesina (2013) investigated the effect of K-promotion on
alumina-supported Mo-Ni catalyst for propane dry reforming in a quartz fixed-
bed reactor at 823973 K and 0.1 MPa. K-promotion increased BET surface area,
pore volume and metal dispersion of the catalyst. The acid:basic site ratio of the
K-containing catalyst was lower than in the undoped Mo-Ni counterpart. The two
catalysts have comparable propane dry reforming performance although CO
formation was greater on the promoted catalyst. It was, however, accompanied by
poorer H
2
production rate. The optimum CO
2
:C
3
H
8
ratio, R
opt
= 2, was below the
stoichiometric value for H
2
on both catalysts but at about 3 for CO production. It
was dound that potassium addition reduced the carbon deposition on the used
catalyst and improved its stability. Post-reaction characterization by double cycle
TPR-TPO runs confirmed the existence of at least two types of carbonaceous
species (atomic and dehydropolymerised, aged) in the used catalysts. It was also
confirmed that the physicochemical properties of the regenerated catalysts could
not be completely restored to the initial state but depended on the reforming
conditions.
Chesterfield et al (2013) performed an evaluative investigation of
propane CO
2
reforming over bimetallic alumina supported 5%Mo10%Ni catalyst
in a fixed-bed reactor at 0.1 MPa and temperatures ranging from 823 to 973 K. In
this investigation, it was found that temperature-programmed calcination revealed
the metal nitrate decomposition of MoNi catalyst at around 475 K, while TPR
analysis showed the complete reduction of fresh catalyst at 973 K. Apart from it,
physicochemical properties of MoNi/Al
2
O
3
catalyst were determined from N
2
-
physisorption, H
2
-chemisorption, NH
3
and CO
2
-temperature-programmed
desorption as well as X-ray diffraction analysis in which NH
3
-TPD analysis
revealed that the MoNi/Al
2
O
3
catalyst surface was populated with weak and
strong Lewis acid sites, while CO
2
-TPD showed only one peak assigned to weak
Lewis basic site. It was also demonstrated that increased CO
2
partial pressure in
feed had caused an initial increase in H
2
, CO, and CO
2
reaction rates, peaked at
stoichiometric feed ratio (R), CO
2
:C
3
H
8
, of 3 followed by a gentle decrease for
R>3. H
2
:CO and CH
4
:CO experienced logarithmic decay with increased reactant
ratio, while CO:CO
2
ratio had remained constant at 2. Also, time-on-stream
analysis revealed that MoNi is a stable and active catalyst at different
temperatures during propane dry reforming. In the end, a global kinetic
expression corresponding to carbon deposition was obtained over various feed
compositions and temperatures. Post-reaction data from TPRTPOTPRTPO
confirmed the presence of at least two carbon pools, C

and C

. The former,
atomic carbon is reactive with H
2
and the latter, dehydropolymerised and aged
carbonaceous matter, is only removed by O
2
.
Althenayan et al. (2010) estimated reaction metrics from longevity runs
by bimetallic CoNi/Al
2
O
3
catalyst for propane dry reforming because dry
reforming of hydrocarbons is accompanied by carbon deposition making it
difficult to unambiguously estimate the true reaction metrics (rate constant, yield
and selectivity) without the masking effect of coke formation. This study
employed a method originally proposed by Levenspiel [1999. Industrial &
Engineering Chemistry Research 38, 41404143] to determine the intrinsic
reaction rate simultaneously with the carbon-induced deactivation coefficient
from transient rate data over an extended period of time (up to 72 h), for propane
dry reforming over a CoNi catalyst at 823973 K. The rate constant k and
deactivation coefficient, k
d
were determined from a fit of the concentration history
data to the hyperbolic reactiondeactivation model for 1st-order kinetics in a plug
flow reactor. However, the product H
2
:CO ratio was generally invariant with time
over the 3-day period for different CO
2
:C
3
H
8
feed ratio values (47) but remained
within a band of between 0.4 and 0.6. Both k and k
d
exhibited a negative order
dependency on the CO
2
:C
3
H
8
ratio at 0.575 and 2.39, respectively. Arrhenius
treatment of these two reaction metrics also yielded activation energy estimates of
92.3 and 164.4 kJ mol1 for the true reforming reaction and deactivation process,
respectively. Catalyst characterization was carried out using XRF (X-ray
fluorescence), liquid N
2
adsorption, XRD (X-ray diffraction), H
2
chemisorption,
temperature programmed desorption of NH
3
and CO
2
, temperature-programmed
reduction (with H
2
) and oxidation (with air) as well as solid TOC content analysis.
Olafsen et al. (2009) compared the CO
2
reforming of methane and
propane over two different Ni catalysts: one reference Ni/SiO
2
system and a Ni/
Mg(Al)O hydrotalcite-derived catalyst, shown previously to display high catalytic
stability for long term reforming. By combining the Tapered Element Oscillating
Microbalance (TEOM), Temperature Programmed Hydrogenation (TPH),
Transmission Electron Microscopy (TEM) and magnetic measurements, the
formation of coke and its role on the catalyst activity had been investigated and
compared for both hydrocarbons. It was found that Ni/SiO
2
and Ni/Mg(Al)O are
both more active for methane reforming than for propane reforming. Coke
formation is much more pronounced for propane than for methane over both
catalysts. However, for both hydrocarbons a much faster carbon formation were
observed over the Ni/SiO
2
catalyst than over the Ni/Mg(Al)O catalyst. The
difference in the rates of coke formation for methane and propane was ascribed in
the case of propane to partially dehydrogenated C
3
ad-species, which proved to be
good coke precursors. The superior stability of the hydrotalcite-derived catalyst
was due to the strong interaction of the nickel phase with the support and the
capacity of the support to activate CO
2
and channel oxygen to the nickel phase.

Jensen et al. (2009) has done mechanistic study of the dry reforming of
propane to synthesis gas over a Ni/Mg(Al)O catalyst. The dry reforming of
propane to synthesis gas over 2 wt.% Ni/Mg(Al)O has been investigated by
means of partial pressure variation experiments, kinetic isotope effect and
isotopic tracer studies, all at 600
0
C. Partial pressure variation experiments
gave reaction orders of 0.18 and 0.36 with respect to propane and carbon
dioxide, respectively, indicating a high surface coverage under the chosen
test conditions (C
3
H
8
:CO
2
:H
2
:N
2
(%) = 10:30:10:50, GHSV (gas hourly
space velocity) = 13333 ml/h gcat). Switching to a deuterated feed did not
induce any kinetic isotope effect, revealing that CH bond rupture is not
involved in the rate-limiting step. Isotopic tracer studies showed that the
reverse water-gas shift reaction approaches equilibrium under the applied
test conditions, demonstrating that CO
2
activation is fast. Temperature-
programmed deuteration of a used catalyst showed that C
3
species
dominated among the hydrocarbon species on the catalyst surface.
Together, the above results led to the conclusion that CC bond rupture was
the rate determining step of reaction. Isotopic tracer studies further showed
(1) that methane is mainly formed as a primary product of reaction by
propane dissociation followed by hydrogenation on Ni sites, and (2) the
primary reaction selectivity favored C oxidation to CO over C
hydrogenation to methane. Co-feeding methane and propane with CO
2

showed that propane is converted significantly faster than methane.
Rberg et al. (2007) tested Ni-based catalysts for propane dry reforming
to synthesis gas to study influence of support and operating parameters on catalyst
activity and stability. A series of supported Ni catalysts (2 wt% Ni) were tested
for the dry reforming of propane to synthesis gas reaction at 600
0
C,
C
3
H
8
:CO
2
:H
2
:N
2
(%) = 10:30:10:50, 1atm and GHSV (gas hourly space velocity)
= 13333 Nml/(h g
cat
). The support systems studied range from acidic (SiO
2
) to
amphoteric [Al
2
O
3
, Mg(Al)O] to basic oxides [MgO, CaO], with BET areas of
100, 100 (131 and 173), 117, and 16 m
2
/g, respectively. Ni particle size and
reduction state was investigated by magnetic measurements, performed before
and during catalytic testing. Support basicity was measured by CO
2
TPD. It was
observed that the intrinsic activity of the catalysts (TOF) depends strongly on the
Ni particle size (D) in the size range of 4.511 nm. As such, a TOF versus D plot
showed an exponential inverse relationship, much more pronounced than
previously reported for the methane dry reforming reaction. In turn, a volcano-
type correlation was found between intrinsic catalyst activity and support basicity.
As a result, the best catalytic performance, including a low selectivity to methane,
was observed on the most dispersed Ni particles supported on Mg(Al)O. Key
mechanistic features were derived from these observations.
Olafsen et al. (2005) studied Mechanistic features for propane reforming by
carbon dioxide over a Ni/Mg(Al)O hydrotalcite-derived catalyst. A 1.9 wt%
Ni/Mg(Al)O hydrotalcite-derived catalyst was studied for the dry reforming of
propane to synthesis gas at 600
0
C and 1 atm. The catalyst showed limited initial
deactivation and then was exceptionally stable throughout a 34-day test. Catalyst
characterisation indicates that the carrier material consists of a mixed Mg(Al)O
phase before and after testing, and that carbonate forms on the support surface
under dry reforming conditions. The Ni particles are in close contact with, and
partially decorated by, the basic support. No carbon whisker formation was
observed by transmission electron microscopy after catalytic testing. Alternating
pulse experiments in a Temporary Analysis of Product-II (TAP-II) reactor system
indicated that CO
2
was associatively adsorbed on the basic Mg(Al)O carrier and
acts as a permanent source of oxygen species for the Ni metal. Propane reacted
rapidly with Ni-O species to form CO and H
2
O. Under TAP conditions, reduced
Ni reacts gradually with carbonate from the support to give Ni-O species and CO.
Solymosi et al. (2003) studied CO
2
reforming of propane over supported
Rh. The adsorption, decomposition, and reaction of propane with CO
2
have been
investigated on Rh catalysts, deposited on various supports. The strong interaction
of propane with Rh was noticed above 273 K. By means of Fourier transform
infrared spectroscopy, -bonded propylene, di--bonded propylene, and
propylidyne have been identified. Propane underwent dehydrogenation and
cracking on supported Rh at 824923 K. Propylene formed with a selectivity of
5060%. The other major products were ethylene and methane. The deposition of
carbonaceous species was also observed, the hydrogenation of which occurred
only above 700750 K, with a peak temperature of 900950 K. The amount of
carbon was more than one order of magnitude higher than that of surface Rh
atoms, suggesting its diffusion from the Rh onto the support. The presence of
CO
2
basically altered the reaction pathway of propane, and the formation of
H
2
and CO with a ratio of 0.420.59 came into prominence. Propylene was
detected only in traces. This led to the assumption that propylene reacted quickly
with CO
2
over Rh after its formation. This idea was confirmed by separate study
of the reaction of propylene with CO
2
. Taking into account the rates of
decomposition of propane and CO
2
on Rh catalysts, as well as the reaction orders,
It had come to the conclusion that the CO
2
was involved in the rate-determining
step of the dry reforming of propane. The highest specific rates for the production
of H
2
and CO were measured for Rh/TiO
2
; this was explained by the extended
dissociation of CO
2
due to the electronic interaction between Rh and TiO
2

_

4.2 THERMODYNAMICALLY STUDIES
Sharma and Schoegl (2013) campared the performance of different
pathways for gas-phase (non-catalytic) fuel reforming between 600 and 1000 C.
Specifically, the conversion of propane to hydrogen-rich syngas had been
investigated numerically and experimentally for pyrolysis (Py), steam reforming
(SR), partial oxidation (POx), and autothermal reforming (ATR). All the
experiments were conducted in a tubular quartz reactor, where temperatures were
imposed externally; reactants were diluted with nitrogen to reduce the impact of
endothermic/exothermic reactions on the variation of gas-phase temperatures. In
experiments, product concentrations of hydrogen, carbon monoxide, carbon
dioxide, methane, and a range of hydrocarbon species were measured at
predetermined operating conditions. The performance of each homogeneous
reforming process was evaluated and compared by assessing propane conversion
and production efficiencies for hydrogen and other species of interest. At 600 C,
propane conversion was low, but increased substantially with temperature;
complete conversion was achieved at 1000 C. Furthermore, findings show
improved hydrogen production efficiencies of POx/ATR when compared to
Py/SR. Experimental results were also substantiated by numerical simulations
with detailed chemical kinetics; numerical results are in good agreement with
experiments at identical operating conditions. Experimental and numerical results
for non-catalytic propane reforming at all tested temperatures (6001000 C)
implies a negligible impact of steam addition to the process, as results for SR
resemble pyrolysis results, and ATR closely follows partial oxidation
characteristics. As such, results clearly show that steam does not play an active
role in gas-phase reforming of propane at intermediate temperatures.
Wang et al. (2010) performed thermodynamic analysis of steam
reforming (SR) and dry reforming (DR) of propane using Gibbs free energy
minimization method. They investigated equilibrium composition as a function of
temperature (700-1100K), pressure (1-5atm), CO
2
/propane ratio (CPR, 1-12) and
H
2
O/propane ratio (WPR, 1-18). It was demonstrated that from a thermodynamic
perspective, for synthesis gas production which is having useful composition for
chemical production; dry reforming is better than steam reforming while for
hydrogen production steam reforming is better than dry reforming. It was shown
that conversion of propane was 100% for the range of variables used in this study.
One atmospheric pressure is preferable for both DR and SR. They also mentioned
different condition for different ratio of H
2
/CO in synthesis gas. For example for
H
2
/CO ratio=1, 975K and CO
2
/propane ratio (CPR) = 3 was found optimum; for
H
2
/CO=2, 1100K and CO
2
/propane ratio (CPR)=3 was found optimum. For
temperature more than 1000K and CPR more than 6, around 10 moles of
synthesis gas was produced per mole of propane. It was also reported optimum
condition for hydrogen generation from steam reforming i.e. temperature 925-
975K and water to propane ratios WPR =12-18. At this condition production of
hydrogen was 9.1 mol per mol of propane and also at that condition methane and
carbon formation is suppressed.

Kale et al (2010) studied the thermodynamic feasibility of a new process
scheme combining chemical looping combustion (CLC) and combined reforming
(CR) of propane (LPG) is studied in this paper as existing energy generation
technologies emits CO
2
gas and are posing a serious problem of global warming
and climate change. In this study of CLC of propane with CaSO
4
as oxygen
carrier is observed in respect to thermodynamic feasibility in temperature range
(400782.95 C) at 1 bar pressure. This showed that the CO
2
generated in the
CLC could be used for combined reforming of propane in an autothermal way
within the temperature range (4001000 C) at 1 bar pressure to generate syngas
of ratio 3.0 (above 600 C) which is extremely desirable for petrochemical
manufacture. The process scheme generated (a) huge thermal energy in CLC that
can be used for various processes, (b) pure N2 and syngas rich streams can be
used for petrochemical manufacture and (c) takes care of the expensive CO2
separation from flue gas stream and CO
2
sequestration. The thermoneutral
temperature (TNP) of 702.12 C yielding maximum syngas of 5.98 mol per mole
propane fed, of syngas ratio 1.73 with negligible methane and carbon formation
was identified as the best condition for the CR reactor operation. The process can
be used for different fuels and oxygen carriers.
Solymosi et al (2005) observed dry reforming of propane over supported
Re catalyst. In this, Fourier-transformed infrared spectroscopy was used for
observing interactions between propane and Re/Al
2
O
3
catalyst. Fourier-
transformed infrared spectroscopy revealed that there is no strong interaction
between propane and Re/Al
2
O
3
catalyst at 250300 K producing di--bonded
propylene or propylidyne whereas, CO
2
was adsorbed mainly molecularly on
supported Re reduced at 673 K, also, the presence of propane induced its
dissociation even at 300 K resulting in the formation of adsorbed CO absorbing at
2041 cm
1
. In addition, the co-adsorption of the two compounds 5% Re/Al
2
O
3
at
373573 K lead to the formation of formate species. Apart from all these,
Re/Al
2
O
3
catalyzed the dehydrogenation and cracking of propane at 773923 K.
The selectivity of propylene formation was found 4374%. It is also observed that
the addition of CO
2
to propane dramatically affected the reaction pathway, and,
instead of the dehydrogenation process, the formation of H
2
and CO with a ratio
of 0.560.61 became the dominant route. Also, the highest conversion values
were measured for the Re/Al
2
O
3
reduced at 673 K. The steady-state conversion of
propane was also found depended on the composition of the reacting gas mixture:
it was 50% at C
3
H
8
/CO
2
(1/3) and 80% at C
3
H
8
/CO
2
(1/6). The deposition of
carbon was observed, the extent of which can be lowered with increasing CO
2

content of the reacting mixture. From the kinetic studies it was inferred that the
CO
2
is involved in the rate-determining step of the dry reforming of propane. As
propylene was not detected or was detected only in traces, it was assumed that the
hydrocarbon fragments formed in the activation of propane reacted quickly with
adsorbed oxygen and CO
2
. A possible mechanism for the dry reforming of
propane on Re catalysts can be proposed from this study.

4.3 MODELING STUDY
Yu et al. (2011) studied non-thermal plasma (NTP) assisted CO
2

reforming of propane in combination with a Ni/-Al
2
O
3
catalyst. The activation
temperature of the catalyst was reduced, and the conversion of the reactants was
improved compared with the summed results of the plasma and catalytic modes.
The selectivity of H
2
and CO was greatly enhanced. Both excited species
produced by NTP and the interaction of NTP with the catalyst were probably
responsible for the decrease of the activation temperature of the catalyst.
Synergistic effect between the plasma and the catalyst lead to the improvement of
the conversion of reactants and the selectivity of H
2
and CO.
Huang et al (2011) set up an anode-supported SOFC unit with NiYSZ
anode operated at 800 C and tested it with direct feeding of 5% propane to
understand the reaction mechanism of propane internal reforming in the solid
oxide fuel cell (SOFC). This mechanism is important for the design and operation
of SOFC internal processing of hydrocarbons. CO
2
reforming of propane was
carried out in a reactor with NiYSZ catalyst to simulate internal propane
processing in SOFC. The performance of this direct propane SOFC was found
stable. He found that the C specie formed over the anode functional layer of
SOFC could be completely removed. The major gas products of SOFC were H
2
,
CO, CH
4
, C
2
H
4
and CO
2
. Pseudo-steady-state internal processing of propane in
the anode catalytic layer of SOFC was associated with a CO
2
/C
3
H
8
molar ratio of
about 1.26 and basically CO
2
reforming of propane. It can be assumed that CO
2

dissociation to produce the O species to oxidize the C species from
dehydrogenation and dissociation of propane and its fragments should have been
the major reaction during CO
2
reforming of propane.

You might also like