Bibi Abrahim MSC Thesis

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 88

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/330293471

Syngas Conversion from Biomass Reforming to Produce Alternative Diesel


Fuel

Thesis · September 2017

CITATIONS READS

0 2,406

2 authors:

Nariefa Abrahim Valerie Dupont


University of Otago University of Leeds, Leeds, United Kingdom
10 PUBLICATIONS 14 CITATIONS 88 PUBLICATIONS 3,487 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Nariefa Abrahim on 10 January 2019.

The user has requested enhancement of the downloaded file.


MSc. Energy and Environment

CAPE5000M Research Project

FINAL REPORT

Syngas Conversion from Biomass Reforming to Produce Alternative


Diesel Fuel for India

Student: Bibi Nariefa Abrahim

Supervisor: Dr. Valerie Dupont

The University of Leeds


School of Chemical and Process Engineering

5 September 2017
Abstract
Energy has been found to be one of the greatest requirement for a country’s social and economic development.
Developing countries, specifically India, can benefit tremendously from the exploration and development of cleaner,
more affordable and easily accessible alternative sources of energy. In India, the transport sector is one of the
greatest contributor to air pollution and a major energy consumer. Diesel fuel, has been found to be the most
demanded fuel type with respect to liquid transport fuel. It is therefore necessary that research and studies be
conducted within the area of alternative diesel fuels. Dimethyl ether is one such fuel that has similar properties as
diesel, but with reduced emissions. Dimethyl ether is a simple ether that can be produced from the catalytic reaction
of synthesis gas, commonly referred to as syngas. Syngas conversion from biomass to Dimethyl ether occurs in
two stages namely methanol synthesis and dehydration. This study investigated the conversion of biomass-derived
syngas to dimethyl ether, using the reaction kinetics of methanol formation and dehydration, in an attempt to
develop to a practical model for the ideal production of an alternative diesel fuel.
In this study the production of dimethyl ether from syngas was based on the two-staged process of methanol
synthesis and dehydration. Aspen Plus V8.8 software was used to develop process flow sheets for each stage of
the process, indicating the main components along the production pathway. The models employed in this study
were based on single -pass catalytic, shell and tube fixed-bed reactors, where isothermal and steady-state
conditions were assumed. The mechanism of the methanol synthesis model was sourced from Graaf et al in 1988,
where the hydrogenation of carbon monoxide, carbon dioxide and the reverse water-gas shift reaction were
included in the kinetics. The kinetics presented by Graaf et al, were applied using Microsoft Excel to derive a
practical model for the conversion of partially-oxidised syngas to methanol, using the hydrogenation of the carbon
oxides and the reverse water-gas shift reactions. Using the reaction rates of the main pathways, the required
catalyst weights and reactor specifications were derived. The reactions of carbon monoxide and carbon dioxide
hydrogenation and the reverse-water gas shift reaction were examined analysed based on the sensitivity to input
parameters such as pressure, temperature and feed-gas composition. It was found that the hydrogenation of
carbon monoxide occurred at the highest rate during majority of the reaction time, and favoured much lower
temperatures than carbon dioxide. The rate of the reverse water-gas shift reaction was found to be negative,
indicating that the reaction occurred in the opposite direction. Further analysis of the synthesis model revealed the
thermodynamic limitations of the chemical reactions, whereby the exothermicity and non-equimolar state led to
favourable conversion rates at lower temperatures and higher pressures. Additionally, the model was tested
against varying feed-gas compositions to determine the effect on the conversion rates of the carbon oxides, where
it was found that a carbon monoxide-rich gas may be favourable for methanol production. The detailed review of
the methanol synthesis process indicated that optimal equilibrium conversion and production of methanol occurred
at 483K and 85 bars, with suitable catalyst requirements, reactor specifications and product composition.
The second stage of the production process was investigated with respect to the dehydration reaction of methanol
to produce dimethyl ether. The kinetics applied for the methanol dehydration process was sourced from Bercic et
al (1993), where the mole fractions of the reaction species along with equilibrium constants were used to derive a

II
reaction rate. The kinetics were tested against varying compositions of feed-gas, where it was revealed that
greatest dimethyl ether was produced with inlet feeds containing the highest methanol concentration. However,
as the methanol levels in the feed -gas increased, so did the catalyst weights and the reactor lengths required for
the conversion process. Additional simulation of the dehydration kinetics revealed that due to the equimolar state
of the dehydration reaction, the conversion and production of the alternative fuel was independent of pressure
change, allowing the system to operate at low pressures. The sensitivity analysis of the model also indicated that
the reaction rates were unaffected by pressure, given the exclusion of partial pressures in the equation. However,
as seen in the synthesis model, the dehydration process was also restricted by thermodynamics, whereby the
exothermic reaction indicated greater methanol conversions at lower temperatures. On the other hand, it was
revealed that the dehydration of methanol at very low temperatures required vast amounts of catalyst, resulting in
the selection of 573K as the ideal temperature for the reaction, given the favourable catalyst and reactor
specifications.
The simulation results for methanol synthesis and dehydration reported in this study produced fairly accurate
results, given the verification with the literature. Each model was validated against past studies, where
approximately 1.0% deviation were reported. The models were extrapolated for increased productions at
commercial level, where favourable results regarding catalyst and reactor requirements were indicated. The kinetic
models for the synthesis and dehydration of methanol to dimethyl ether produced in this study, can be therefore
be applied to simulate the conversion of biomass-derived syngas in a catalytic, shell and tube fixed-bed reactor to
produce an alternative diesel fuel for India.

III
Acknowledgements
The researcher would like to express deep gratitude to Allah for his guidance throughout this report. Many thanks
are also extended to my project supervisor, Dr Valerie Dupont for her continuous guidance, comments and critical
inputs at every stage of the research. The time, dedication and valuable reviews from Dr Dupont are greatly
appreciated. I would also like to thank Mr Robert White, for his time, guidance and valuable contribution towards
this project.
The ESPRC is acknowledged for the GCRF grant EP/P51097X/1 (Integrated Low Carbon Energy Solutions for
Remote Rural Area Project).
Finally, I would like to express gratitude to my family and friends for the prayers, encouragement and motivation
throughout this study, as your support enabled the successful completion of this research.
Contents
Abstract ................................................................................................................................................................. II

Acknowledgements .............................................................................................................................................IV

Chapter 1: Introduction ........................................................................................................................................ 1

1.1 Background ................................................................................................................................................. 1


1.1.1 Global Warming and Fossil Fuels.......................................................................................................... 1
1.1.2 Energy Utilization and India ................................................................................................................... 2
1.1.3 Alternative Energy Sources ................................................................................................................... 4
1.1.4 Dimethyl Ether: Substitutability for India’s Diesel Fuel .......................................................................... 5
1.2 Research Context ....................................................................................................................................... 6
1.3 Project Scope .............................................................................................................................................. 8
1.4 Aim of Study................................................................................................................................................ 9
Chapter 2: Literature Survey- DME Production ................................................................................................ 10

2.1 Introduction............................................................................................................................................... 10
2.2 DME Production Techniques ................................................................................................................... 10
2.2.1 Fischer-Tropsch Process .................................................................................................................... 10
2.2.2 JFE Direct DME Synthesis Process .................................................................................................... 11
2.2.3 Two-Stage Production of DME ............................................................................................................ 13
2.2.4 Direct vs Indirect Production of DME................................................................................................... 14
2.3 Reactor Models ......................................................................................................................................... 15
2.4 Catalysts Selection ................................................................................................................................... 17
2.5 Effects of Operating Conditions on DME Production ........................................................................... 18
2.5.1 Methanol Synthesis: Effects of Temperature and Pressure ................................................................ 18
2.5.2 Methanol Dehydration: Effects of Temperature and Pressure............................................................. 20
2.5.3 Isothermal Conditions for Methanol Synthesis and Dehydration ......................................................... 20
2.6 Kinetic Models for Methanol Synthesis .................................................................................................. 21
2.7 Kinetic Models for Methanol Dehydration .............................................................................................. 25
2.8 Conclusion ................................................................................................................................................ 26
Chapter 3: Methodology ..................................................................................................................................... 27

3.1 Introduction............................................................................................................................................... 27
3.2 Methanol Synthesis .................................................................................................................................. 27
3.2.1 Process Pathway................................................................................................................................. 27
3.2.2 Feed-Gas Composition ....................................................................................................................... 27
3.2.3 Kinetics of Methanol Synthesis............................................................................................................ 28
3.3 Methanol Dehydration .............................................................................................................................. 32
3.3.1 Process Flow ....................................................................................................................................... 32
3.3.2 Feed-Gas Composition ....................................................................................................................... 32
3.3.3 Kinetics of Methanol Dehydration ........................................................................................................ 32
3.4 Reactor Specifications and Flow Properties .......................................................................................... 34
3.5 Sensitivity Analyses: Effects of Temperature, Pressure and Feed-Gas Composition ....................... 35
3.6 Conclusion ................................................................................................................................................ 36
Chapter 4: Results & Discussion....................................................................................................................... 37

4.1 Methanol Synthesis .................................................................................................................................. 37


4.1.1 Process Flow of Methanol Synthesis................................................................................................... 37
4.1.2 Methanol Synthesis Model .................................................................................................................. 37
4.1.3 Sensitivity Analyses of Methanol Synthesis ........................................................................................ 42
4.1.4 Methanol Synthesis Model Predictions: Increased Production Level .................................................. 48
4.1.5 Validation of Methanol Synthesis Model.............................................................................................. 49
4.2 Methanol Dehydration .............................................................................................................................. 49
4.2.1 Process Flow of Methanol Dehydration ............................................................................................... 49
4.2.2 Methanol Dehydration Model............................................................................................................... 50
4.2.3 Sensitivity Analyses of Methanol Dehydration..................................................................................... 54
4.2.4 Dehydration Model Predictions: Increased Production Level .............................................................. 58
4.2.5 Validation of Methanol Dehydration Model .......................................................................................... 58
4.3 Conclusion ................................................................................................................................................ 59
Chapter 5: Conclusion & Recommendations for Further Work ...................................................................... 60

5.1 Conclusion ................................................................................................................................................ 60


5.2 Future Work............................................................................................................................................... 61
References........................................................................................................................................................... 62

List of Figures
Figure 1.1: Global GHG Concentrations for 1850-2012 (IPCC, 2014) .................................................................... 1
Figure 1.2: Global Temperatures for the Period of 1880-2000 (National Aeronautics and Space Administration,
2016) ....................................................................................................................................................................... 2
Figure 1.3: Energy Utilization in India in kg of Oil Equivalent per Capita (World Bank, 2017) ................................ 3
Figure 1.4: Projected Energy Use by Vehicle Type in India in Petajoules (PJ), (Adapted from European Business
and Technology Centre, 2013)................................................................................................................................ 3
Figure 1.5: Current and Projected CO2 Emissions for India (International Energy Agency, 2011) .......................... 4
Figure 1.6: Line Plot of India’s Current and Projected Biofuel Consumption Patterns (World Energy Council,
2011) ....................................................................................................................................................................... 4
Figure 1.7: Available Quantities of Plant-based Residues in India (Ghosh, 2016) .................................................. 6

VI
Figure 1.8: DME Production Plants and Production Capacities (Taupy, n.d) .......................................................... 7
Figure 1.9: Possible Production Pathway for DME with the Outlined Scope of this Study ...................................... 8
Figure 2.1: FT Synthesis Process to Produce DME/Diesel-based Fuel (Puladian et al., 2013) ............................ 11
Figure 2.2: JFE Direct DME Production Plant with Production Flowchart (Bourg, 2006) ...................................... 12
Figure 2.3: Equilibrium Conversion of Methanol and DME at 50 bars and 533K (Ohno et al., 2005) ................... 13
Figure 2.4: Reaction Pathways for CO and CO2 to CH3OH (Yang et al., 2013) .................................................. 14
Figure 2.5: Schematic of a Slurry, Fluidised Bed and Fixed Bed Reactors (Laurence, 2009) .............................. 16
Figure 2.6: Effects of the Catalyst Performance and Activities with Varying Amounts of Al2O3 (Xiao et al., 2017)18
Figure 2.7: Conversions at Various Temperatures (Arthur, 2010) ........................................................................ 19
Figure 2.8: Mole Fractions of Reactions Species along Reactor, along with the Effect of Pressure on Methanol
Formation (Mayra et al., 2008)Conversions at Various Temperatures (Arthur, 2010) .......................................... 19
Figure 2.9: CO Conversion against Reaction Temperature in an Isothermal System (Alam, 2011) ..................... 21
Figure 2.10: Mole Fractions of DME in Adiabatic and Isothermal Systems (Farsi, 2015) ..................................... 21
Figure 2.11: Mole Fractions of DME in Adiabatic and Isothermal Systems (Farsi, 2015) ..................................... 21
Figure 4.1: Hypothetical Process Flowsheet of the Methanol Synthesis Process ................................................. 37
Figure 4.2: Conversions of CO and CO2 along the Length of Reactor .................................................................. 39
Figure 4.3: Rates of the Reaction along the Reactor ............................................................................................ 40
Figure 4.4: Reaction Rates versus Weight of Catalyst .......................................................................................... 41
Figure 4.5: Mole Fraction of Reaction Species along the Length of Reactor ........................................................ 42
Figure 4.6: Mole Fractions of Methanol at Varying Pressures .............................................................................. 43
Figure 4.7: Mole Fractions of Methanol at Varying Temperatures ........................................................................ 45
Figure 4.8: CO Conversion Rates along the Reactor for Varying Feed-Gas Compositions .................................. 46
Figure 4.9: CO2 Conversion along the Reactor for Varying Feed-Gas Compositions ........................................... 47
Figure 4.10: Weight of Catalyst along the Reactor Length for Varying Feed-Gas Compositions .......................... 48
Figure 4.11: Proposed Process Flow Sheet for Methanol Dehydration................................................................. 50
Figure 4.12: Conversion of Methanol along the Length of Reactor ....................................................................... 52
Figure 4.13: Reaction Rate of Methanol Dehydration across the Length of Reactor ............................................ 53
Figure 4.14: Reaction Rate vs Weight of the Catalyst for Methanol Dehydration ................................................. 53
Figure 4.15: Mole Fractions of the Main Reaction Species along the Reactor...................................................... 54
Figure 4.16: Mole Fractions of DME at Varying Temperatures ............................................................................. 56
Figure 4.17: Conversion of Methanol for Varying Feed-Gas Compositions .......................................................... 57

List of Tables
Table 1.1: Population and Gross Domestic Product (GDP) Growth (%) for India (World Energy Council, 2011) ... 2
Table 1.2: Properties of Diesel, Biodiesel and DME (Roh et al., 2015) ................................................................... 5
Table 2.1: Operating Conditions for the Direct and Indirection Conversion of DME (Ohno et al., 2005)............... 14
Table 2.2: Properties of the Three Main Reactor Types used for DME Synthesis (Hu et al., 2011) ..................... 15

VII
Table 2.3: Initial Kinetic Models for Methanol Synthesis ....................................................................................... 23
Table 2.4: Additional Kinetic Models for Methanol Synthesis................................................................................ 24
Table 2.5: Kinetics Models for Methanol Dehydration Presented in Literature...................................................... 26
Table 3.1: Syngas Composition used in the Study................................................................................................ 28
Table 3.2: Reaction Rate Constants for Methanol Synthesis ................................................................................ 29
Table 3.3: Adsorption Equilibrium Constants for Methanol Synthesis ................................................................... 30
Table 3.4: Reaction Equilibrium Constants for Methanol Synthesis ...................................................................... 30
Table 3.5: NASA CEA Equilibrium Conversion of CO and CO2 ............................................................................ 31
Table 3.6: Rates of Change in Flow Rates and Respective Flow Rates of Reaction Species along the Reactor . 31
Table 3.7: Rate of Change in Weight of Catalyst .................................................................................................. 31
Table 3.8: Conversion Rates of CO and CO2 as Predicted by the Model ............................................................. 32
Table 3.9: Kinetic Constants used in the Methanol Dehydration Model ................................................................ 33
Table 3.10: Kinetic Parameters used in the Selected Model................................................................................. 33
Table 3.11: Mole Fraction and Partial Pressure Relationships ............................................................................. 33
Table 3.12: Rate of Change in Molar Flow Rates and the Respective Flow Rates of Reaction Species .............. 34
Table 3.13: NASA CEA Equilibrium Conversion of Methanol ............................................................................... 34
Table 3.14: Rate of Change in Weight of Catalyst and Total Catalyst Weight along the Reactor Length ............. 34
Table 3.15: Catalyst Densities of Methanol Synthesis and Dehydration Catalysts ............................................... 35
Table 4.1: Hypothetical Process Flowsheet of the Methanol Synthesis Process .................................................. 35
Table 4.2: Composition of Syngas and Outlet Mole Fractions as Presented by NASA CEA and this Model ........ 35
Table 4.3: Conversion Rates of CO and CO2 ....................................................................................................... 35
Table 4.4: Effects of Pressure on Conversion Rates, Catalyst Weights and Reactor Lengths ........................... 435
Table 4.5: Conversion Rates, Catalyst Weights and Reactor Lengths at Varying Temperatures ......................... 44
Table 4.6: Cases 1-4 with the Respective Compositions ...................................................................................... 46
Table 4.7: Properties of the Reactor Model for Increased Methanol Production ................................................... 49
Table 4.8: Model Predictions Compared to Literature Data .................................................................................. 49
Table 4.9: Properties of the Recommended Methanol Dehydration Model ........................................................... 51
Table 4.10: Composition of Inlet Gas and Outlet Mole Fractions as Presented by NASA CEA and the Model .... 51
Table 4.11: Methanol Conversion Rates Produced by NASA CEA and this Model .............................................. 52
Table 4.12: Conversion Rates, Catalyst Weights and Reactor Lengths at Varying Temperatures ....................... 55
Table 4.13: Cases 1-3 with the Respective Compositions .................................................................................... 57
Table 4.14: Properties of the Reactor Model for Increased DME Production........................................................ 58
Table 4.15: Comparison of Model Predictions with Literature Data ...................................................................... 59

VIII
Nomenclature
A: Area of Reactor (m2)
Al: Aluminium
Al2O3: Alumina
-Al2O3: Gamma Alumina Catalyst
C: Carbon
CH3OH: Methanol
CH3OCH3: Dimethyl Ether
CH4: Methane
CO2: Carbon Dioxide
CO: Carbon Monoxide
Cu: Copper
Cu/ZnO: Copper-Zinc Oxide
Cu/ZnO/Al2O3: Copper- Zinc Oxide Alumina
dFi: Change in Molar Flow Rate of Component ‘i’ (mols-1)
dW: Change in Weight of Catalyst (kg)
ƒi: Fugacity of Component ‘i’ (bar)
F: Molar Flow Rate (mols-1)
Fi: Molar Flow Rate of Component ‘i’ (mols-1)
Fi,in: Inlet Molar Flow Rate of Component ‘i’ (mols-1)
Fi, enditer: Molar Flow rate of Component ‘i’ at End of Iteration (mols-1)
Fi,eqm: Molar Flow of Component ‘i’ after Equilibrium Conversion (mols-1)
Fi, previous iter: Molar Flow Rate of Component ‘i’ in the Previous Iteration (mols-1)
∆H: Change in Enthalpy (kJmol-1)
H+: Hydrogen ion
H2: Hydrogen
H2COO: Dioxirane
HCO2: Formate
H2O: Water
Ki: Adsorption Equilibrium Constant (bar-1)
ki: Reaction Rate Constant (molkg-1s-1)
Keqm: Thermodynamic Equilibrium Constant
KPi: Equilibrium Constant Based on Partial Pressures
K: Kelvin
kg: Kilogram
L: Length of Reactor (m)

IX
m: Metre
NH3: Ammonia
N2O: Nitrous Oxide
NOx: Nitrogen Oxides
O2: Oxygen
OH: Hydroxide
P: Pressure (bar)
Pi: Partial Pressure of Component ‘i’ (bars)
PT: Total Pressure (bars)
PM: Particulate Matter
Q: Volumetric Flow rate (m3s-1)
R: Ideal Gas Constant (Jmol-1K-1)
ri: Rate of Reaction ‘i’ (molkg-1s-1)
SBET: Surface Area (m2)
SO2: Sulphur Dioxide
T: Temperature (K)
Vol %: Volume Percent
v: Velocity (ms-1)
W: Weight of Catalyst (kg)
Wt. %: Weight Percent
Υi: Mole Fraction of Component ‘i’
Υi,in: Inlet Mole Fraction of Component ‘i’
Yi,eqm: Mole Fraction of Component ‘i’ after Equilibrium Conversion
Zn: Zinc
ZnO: Zinc Oxide
Greek Letters
ρ: Density of Catalyst (kgm-3)
List of Abbreviations
DME: Dimethyl Ether
FT: Fischer Tropsch
GBP: Great British Pound
GHGs: Greenhouse Gases
GDP: Gross Domestic Product
HTFT: High Temperature Fischer Tropsch
IEA: International Energy Agency
IPCC: Intergovernmental Panel on Climate Change
LPG: Liquefied Petroleum Gas

X
LTFT: Low Temperature Fischer Tropsch
MTOE: Million Tonnes of Oil Equivalent
NASA: National Aeronautics and Space Administration
PJ: Petajoules
RWGS: Reverse Water-Gas Shift Reaction

XI
Chapter 1: Introduction
This chapter covers a brief outline of the energy situation globally and nationally with respect to India, more
specifically with focus on the transport sector. Dimethyl ether (DME) was found to be a clean and valuable
alternative to the widely-consumed diesel fuel in India. The application of syngas derived from biomass to DME
has extensively been studied and applied in many countries. The issues attached to the production of DME from
syngas are discussed and the objectives of this research are presented in this chapter.
1.1 Background
1.1.1 Global Warming and Fossil Fuels
Over the years, the temperatures and climate of the planet has seen intense changes, some of which has been
linked to anthropogenic activities. The absorption and loss of radiation from the sun by earth and the surrounding
atmosphere contributes to the climate. The atmosphere serves as an absorber of radiation, warming the earth’s
surface with the retained heat energy. Some of this energy, is re-radiated back into space, making the planet
habitable. The circulation of the energy is referred to as the global energy balance, a phenomenon that is greatly
affected by the atmosphere. Gases such as carbon dioxide (CO2), methane (CH4), and nitrous oxide (N2O) are
referred to as greenhouse gases (GHG), given their abilities to absorb heat energy that would otherwise be re-
radiated back into the space (Hansen et al.,2013).
GHGs are released in vast quantities when fossil-based fuels are combusted to satisfy the energy demands of a
growing population. In the 1850’s, fossil fuel became the main source of energy, which ultimately triggered
increased GHG emissions. As time progressed, the demands of the population became even more significant, and
the use of fossil fuel was even more prevalent, a pattern that led to increased concentrations of CH 4, CO2 and
other gases (IPCC Report (2014). The figure below illustrates the global GHG concentrations for the period of
1850-2012, confirming that the presence of these gases in the atmosphere has grown significantly in past decades.

Figure 1.1. Global GHG Concentrations for 1850-2012 (IPCC, 2014)


The prevalence of these gases, coupled with human activities such as deforestation, has been reported to have
significant impact on the global energy balance, where ocean and atmospheric temperatures have risen (Akpan et
al.,2012). Figure 1.2 displays the trend of the global temperatures between a time period of 1880 to 2000, clearly
visible is the increasing trend with time (National Aeronautics and Space Administration, 2016).

1
Figure 1.2. Global Temperatures for the Period of 1880-2000 (National Aeronautics and Space Administration,
2016)

Increased global temperatures has been reported to be closely linked to the melting of ice-caps, sea-level rise and
other impacts on the human and natural surroundings (IPCC, 2014).
1.1.2 Energy Utilization and India
The interconnections of the energy sector with the basic needs of the population has triggered the recognition that
energy is vital for human existence. In many developing nations, the access to basic electrification is still a major
concern, where it has been estimated that approximately 1.3 billion people have no access to electricity with only
1 billion having limited access (Scott et al., 2013). Much of these countries are concentrated in the regions of the
South Asia and Africa, where only about thirty percent of the population can source electricity (Scott et al., 2013).
Given the limited access to energy, various organizations and countries have since developed strategies and
policies to tackle the issues of energy supply and security, with the aim of improving the economic standards of
the countries; however, the degree of penetration of these very policies have seen varying outcomes with
approximately 1.4 billion people are estimated to still lack access to electrical power in 2030 (International Energy
Agency, 2011). The right to clean, safe, and affordable energy is the basic right of every human, a goal that is now
stipulated in the United Nation Sustainable Development Plan (United Nations Development Programme, 2016).
With this in mind, the fight for energy security has now become one of the greatest global concern with much focus
on low income developing countries. One specific point of interest is India, a highly developing economy, ever-
growing population, diverse rural communities and a wealth of natural resources. According to the World Energy
Council (2011), the energy desires of a country is closely linked to its economic growth patterns and demography.
A closer look at India’s economic growth as shown in table 1.1 revealed that the country will see improvement, at
least up to 2020, surprisingly with a slowly decreasing percentage growth in population.
Table 1.1. Population and Gross Domestic Product (GDP) Growth for India (World Energy Council, 2011)

Feature 1980-1990 1990-2008 2008-2020 2020-2035


Growth in GDP (%) 5.6 6.4 7.4 5.6
Population Growth (%) 2.1 1.6 1.2 0.7

2
The growth in the GDP, indirectly infers a greater energy utilization pattern, much of which is sourced from coal,
oil and gas. The figure below taken from World Bank (2017) illustrates the energy usage in India in kg of oil
equivalent per capita for the period of 1974-2014. It is quite visible that as time progressed, the energy sector in
India grew significantly, with the drivers of economic and population growth.

Figure 1.3. Energy Utilization in India in kg of Oil Equivalent per Capita (World Bank, 2017)

By 2012, India became the fourth largest consumer of energy globally, where much of the reliance is on oil usage,
with a growth from 43% in 1990 to approximately 71% in 2012 (Bandyopadhyay et al., 2015). In 2005, it was
reported that the transport sector drained 16% of the total energy in the country, with 12% stemming from primary
energy sources (European Business and Technology Centre, 2013). The European Business and Technology
Centre (2013), has projected that the sector will consume at least 21% of the total energy and possibly 16% of
primary energy by 2020.
A close examination of the consumption patterns of the transport sector in 2004 revealed that 90% of the energy
used, was diesel and gasoline, where 66% of that total was diesel-based (de la Rue du Can et al., 2009). According
to the International Energy Agency (2015), the demand for transport fuels would amount to approximately 280
million tonnes of oil equivalent (MTOE) by 2040, much of which will be drained by diesel-powered systems. Figure
1.4 illustrates the energy consumption of the transport sector by vehicle type for 2005, and the estimated levels by
2020.

Figure 1.4. Projected Energy Use by Vehicle Type in India in Petajoules (PJ), (Adapted from European Business
and Technology Centre, 2013).

3
The figure clearly depicts that much of the projected energy demands of the sector will be connected to the increase
in diesel-powered transport systems namely heavy and light duty trucks and buses. Two major concerns related
to this pattern are price increases and GHG emissions. It has been recognized that the price of diesel has increased
sharply over the years, leading to various issues of affordability and energy security (de la Rue du Can et al., 2009).
With respect to GHG emissions, it has been indicated that a compounded annual growth rate of 4.5% was reported
for the emissions resulting from the transport sector (Ministry of Statistics and Programme Implementation, 2015).
The figure below depicts the CO2 emissions for India, where the transport sector contributes significantly to the
total atmospheric concentrations. The International Energy Agency (2011), also reported projections for the future
with respect to sectoral contributions to CO2 concentrations, which is also represented in Figure 1.5.

Figure 1.5. Current and Projected CO2 Emissions for India (International Energy Agency, 2011)

1.1.3 Alternative Energy Sources


Given the rising demand of the transport sector in India, coupled with energy security and air quality issues, the
development and exploration of alternative fuels including biofuels, became a priority for the nation. Biofuels, are
biologically derived fuels emanating from plant and animal materials processed to produce gaseous, liquid or solid
energy supplies. The production of biofuels in India is relatively in its initial stage, where small scale biogas
digesters, bioethanol and biodiesel plants have been installed (Basavaraj et al., 2012). However, the World Energy
Council (2011) have reported that the growth of the biofuel industry in India seems promising, given its agricultural
background. In figure 1.6, the rising consumption of biofuel in the country is clearly outlined, inferring the potentials
of the industry.

Figure 1.6. Line Plot of India’s Current and Projected Biofuel Consumption Patterns (World Energy Council,
2011)

4
Given the high consumption patterns of diesel fuel, it is imperative that India partakes in the assessment and
evaluation of diesel alternatives. Biodiesel is one such technique, however the biofuel has various uncertainties
connected with its operations, including release of ultrafine particulate matter (PM), and the variations in studies
relating to nitrogen oxides (NOx) emissions (Bansal et al., 2013). A very promising technology is the utilization of
biomass to produce syngas and further DME, a low polluting and economically favourable substitute for diesel.
1.1.4 Dimethyl Ether: Substitutability for India’s Diesel Fuel
DME, also referred to as methyl ether and dimethyl oxide is a colourless, flammable vapour or liquid with a distinct
odour and a molecular formula of CH3OCH3 (Szybist et al., 2014). The fuel is a synthetically derived ether, with a
similar chemical structure as ethanol (Albrecht, 2004). The alternative fuel shares many similarities with diesel;
however, DME has been reported to have a better ignitability than diesel, given its high cetane number and low
auto-ignition temperature (Albrecht, 2004). Additionally, as compared to diesel and biodiesel, DME contains greater
amounts of oxygen resulting in enhanced air-fuel mixing and a cleaner combustion (refer to table 1.2). On the down
side, the fuel has a lower energy density than diesel and biodiesel, requiring greater consumptions, added to this
is its low viscosity and lubricity, factors which greatly affect leakage and deterioration of fuel injection systems and
pumps (Maji et al., 2015). A comparison of the three fuels including diesel, biodiesel and DME is presented in the
table below:
Table 1.2. Properties of Diesel, Biodiesel and DME (Roh et al., 2015)

Property Diesel Biodiesel DME


Cetane No. 40-55 48-65 >55
Auto-Ignition Temperature (K) 523 ---- 508
Lower Heating Value (MJ/kg) 42.5 39.17 27.6
Boiling Point (K) 453-613 588-623 248.5
Flashpoint (K) 333-353 373-443 231
Vapour Pressure at 293K (bars) <0.1 ---- 5.3
Kinematic Viscosity (cSt) 3 ---- <0.1
Oxygen Content (Wt %) 0 11.2 34.8
Stoichiometric Air/Fuel Ratio 14.6 ---- 9

The application of DME in diesel-powered vehicles is a favourable renewable alternative, given its properties and
abilities of substitutability with diesel in compression ignition engines. Due to the concerns of low lubricity, the
utilization of DME may require slight modifications in the fuel pump and injection systems, similar to that used for
liquefied petroleum gas (LPG) (Szybist et al., 2014). Researchers have also indicated that the use of additives,
including the addition of diesel, may enhance its lubricity and can therefore limit the structural modifications
(European Biofuels Technology Platform, 2016). DME is seen as an appropriate diesel alternative, given its
environmental and energy performance especially for countries like India with vast availability of potential
feedstocks.

5
1.2 Research Context
The criticality of energy security, climate change and economic development have led India to the recognition that
renewable, sustainable, affordable and reliable energy sources are crucial research focuses. India, being one of
the greatest populated nations with a fast-growing economy and rising diesel fuel demands in the transport sector,
may be at great risk to volatilities and insecurities in the energy industry. Unfortunately, the rising economic growth
has had minimal effects on the access of energy resources in many parts of the country, where majority of the
population remains energy poor. The economic growth is concentrated around the major cities, where energy
utilization follows closely. This increased trend of energy use has triggered concerns over the impacts on air
pollution and climate change, adding to the pressures of attaining the sustainable development goals. The country
has therefore shifted much of its attention to the development of alternative energy, with research aimed towards
the utilization of readily available and inexpensive indigenous resources, including agricultural wastes.
The agricultural sector of the country is the spine of its economy, providing a source of income for more than 60%
of the population (Nahar et al., 2017). The biomass potential therefore, is vast, with approximately 686 Mt/yr of
gross biomass coming from plant residues including rice straw, coconut shells and husk and other plant-based
wastes (Nahar et al., 2017). The level of plant residues has been projected to increase in the future specifically for
rice, wheat, maize and sugarcane, indicating the availability of potential biofuel feedstocks (refer to figure 1.7). The
figure below illustrates the current and projected levels of plant-based wastes across India:

Figure 1.7. Available Quantities of Plant-based Residues in India (Ghosh, 2016)

Coupled with this, is approximately 109 Mt/yr of biomass emanating from animal waste such as manure, a quantity
that is also projected to increase with time (Nahar et al., 2017). According to Nahar et al. (2017), animal waste, if
left untreated and exposed can result in the release of 55-65 vol.% of CH4, a highly damaging greenhouse gas
with a global warming potential of 21. Biomass, currently contributes to 30% of energy supply, with much

6
exploitation at domestic levels across the rural communities in India; however, when compared to availability of
feedstock, this utilization level is minute, indicating the possibility for further research.
The application of biological waste for energy harnessing, specifically biomass reforming to produce syngas and
possibly DME, is one such research area, which may result in improved access to diesel energy with reduced GHG
emissions. DME is an isomer of the ethanol compound, with no carbon-carbon bonds (Szybist et al., 2014). The
compound has been recognized as an oxygenated fuel with zero particulate and sulphur emissions and low NO X
releases (Wyoming, 2014). Additionally, its ability to quickly decompose in the troposphere, infers that it does not
contribute to bad ozone formation (Laurence, 2009). Given its properties and similarities with diesel fuel, DME may
be a thermodynamically and sustainable energy source for the ever-growing demands of the transport sector in
India. The fuel is produced in many countries including China, Japan, South Korea and Germany and has been
promoted as a cost-effective, sustainable and renewable diesel alternative (refer to figure 1.8). Automobile
manufacturers such as Nissan, Hino, Isuzu, and Mitsubishi are actively involved in the research and development
of DME-powered vehicles. Figure 1.8 depicts some of the most noticeable DME production plants and the
production capacities, indicating the development of DME in the fuel industry:

Figure 1.8. DME Production Plants and Production Capacities (Taupy, n.d)

According to the Bourg (2006) at the 23rd World Gas Conference in Amsterdam, 2006, India has great potential for
DME production, given the accessibility of feedstock and demand for diesel alternatives. Unfortunately, the country
currently has no syngas production plants and very limited research has been conducted in the field, augmenting
the need for related studies. Globally, syngas conversion to DME is an emerging technique investigated by various
researchers in the areas of reactor selection, operating conditions and DME selectivity factors. The conversion
rate of the fuel is closely connected to the process operations, inlet gas compositions, reactor type, catalysts and
reaction rates, inferring that each application of the methodology is different. The production process is highly
sensitive and requires a final product of at least 99.8% DME purity level to be substituted in diesel-powered vehicles
with little adjustments (Roh et al., 2015). This study, is therefore aimed at assessing and developing an ideal gas
reactor model for the conversion of biomass-derived syngas to DME, as an alternative to diesel fuel for India.

7
1.3 Project Scope
The energy consumption patterns in India have been studied by many researchers, with projections of possible
increased dependence on fossil fuels. This dependency has been projected to grow annually, allowing India to
become vulnerable to issues of energy security. Alternative fuels for diesel vehicles, based on indigenous and
renewable resources is therefore crucial to the country’s development. The identification and examination of the
extent of demand for renewable alternatives to diesel fuel in India was the initial objective of this project.
Having established the demand for diesel alterative, an appropriate, low polluting substitute fuel, DME, was
investigated with emphasis on substitutability, potential feedstocks and production pathways. Moreover, specific
focus was aimed towards the examination of past studies relating to parameters such as operating temperature,
pressure, catalysts and reactor type and the possible effects on product composition. The variations in these
parameters allows for the selectivity of the final products, in this scenario a 99.8% DME was recommended. A
possible production pathway for DME stems from the generation of methanol from syngas, followed by dehydration,
each process enhanced by the application of catalysts. This study was steered towards the examination of the
conversion of syngas to DME, based on the reactions and kinetics of methanol formation and dehydration. The
reactor feed gas was based on possible syngas compositions derived from anaerobically digested and partially
oxidised biomass materials that are extensively available in the rural and urban communities across India. Possible
reaction kinetics were sourced from past studies, whereby reactor models were tested in an attempt to determine
appropriate reaction rate equations for both methanol synthesis and dehydration. The equilibrium conversions for
CO, CO2 and methanol (CH3OH) were identified, so that the conversion levels of the main reactor species were
quantified for both processes. Having established the relative reaction kinetics, various specifications including
reactor size and catalyst weight were also determined. The reactor models were then subjected to sensitivity
analyses, whereby, temperatures and pressures were varied and compared to the effect on favourable product
composition. The delineation of this research is illustrated in the figure below, indicating a possible production
pathway for DME, starting from agricultural waste to the final product, whereby the demarcated red section provides
an outline for this study:

Figure 1.9. Possible Production Pathway for DME with the Outlined Scope of this Study

8
This thesis encompasses the conversion of syngas to DME in a two-staged production route with the kinetic
modelling, reaction kinetics and simulation analyses as the scope of the research work, with the attached aim and
objectives discussed in the following section.
1.4 Aim of Study
The general aim of this study was to develop an ideal gas reactor model for the conversion of biomass-derived
syngas to DME, as an alternative to diesel fuel for India.
Research Objectives
The research objectives of this study were:
1. To determine the level of demand for renewable alternatives to diesel fuel in India.
2. To provide a practical design of an ideal gas reactor for the conversion of biomass-derived syngas to
DME.
3. To conduct sensitivity analyses of the reactor models for ideal equilibrium production of DME.
Outline of Report
In order to achieve the objectives of this study, specifically the production of DME from syngas via methanol
synthesis and dehydration, the research work was divided into sections including literature reviews, modelling and
simulation. Chapter 2 address includes a discussion of DME production techniques presented in literature.
Literature focused on mathematical modelling, reaction kinetics and methodologies used for methanol production
and dehydration are reviewed in the chapter.
Chapter 3 contains the methodologies that were employed during in this study, including mathematical models
used for the two-stage production of DME, both of which were validated experimentally, in past research. The
application of software programmes such as Microsoft excel and Aspen Plus to derive reactor models are also
outlined in the chapter.
In Chapter 4, the results and analyses of data are presented and discussed.
The final Chapter, 5, contains the conclusion and future recommendations of this research.

9
Chapter 2: Literature Survey- DME Production
With the growing concerns of fossil fuel depletion and climate change, much research has been directed towards
the exploration of alternative transport fuels such as DME. This chapter entails an examination of the related
studies focused on DME production, with emphasis on reactor type, catalysts selection and reaction kinetics. The
operating conditions of reactors were also analysed with respect to the effect on product composition in order to
determine the ideal conditions for equilibrium DME production.
2.1 Introduction
The production of DME has almost doubled from 100,000-150,000 tonnes in 1990 to 200,000-300,000 tonnes in
the 2000, where China was noted as one of the greatest producer and consumer of the fuel (Baskaran, 2015). In
2010, Sweden engaged in a demonstration project to produce DME from black liquor-derived syngas, which
became the first bio-DME plant in the world (European Biofuels Technology Platform, 2016). DME has been
reported to have lower climate impact as compared to biodiesel, fossil-based diesel, methanol, ethanol and biogas,
which greatly aids in the favourable promotion of the fuel for the transport sector (Kittleson et al., 2010). DME can
be produced from biomass-reformed syngas in a single phase or double phase synthesis process, where syngas
is converted to methanol in the presence of a catalyst, followed by catalytic dehydration. In this chapter the various
DME production techniques are reviewed and discussed, where specific focus is aimed at the two-stage indirect
pathway. In the production process of the fuel, four main chemical reactions occur, specifically hydrogenation of
carbon oxides, reverse water-shift gas reaction and methanol dehydration. These chemical reactions are enhanced
by the application of favourable catalysts, which are employed at specific operating conditions to improve the
conversion processes. In the following sections, the related theories and studies for DME production are examined.
2.2 DME Production Techniques
2.2.1 Fischer-Tropsch Process
The production of liquid biofuels has been around since the early 1900’s, where the Fischer-Tropsch process
introduced by Hans Fischer and Franz Tropsch was employed. Since the development of the DME sector, the
Fischer-Tropsch was adjusted to produce the fuel through the catalytic conversion of gasified-biomass to produce
bio-syngas followed by DME production. Syngas contains varying amounts of hydrogen (H2), CO, carbon dioxide
(CO2), and small quantities of other gases (Xie et al., 2015). The figure below represents the FT process pathway,
from the introduction of biomass to the production of diesel-based fuel:

10
Figure 2.1. FT Synthesis Process to Produce DME/Diesel-based Fuel (Puladian et al., 2013)

The FT process is rather complex, with many possible reaction routes, however some of the main chemical
reactions are presented below (Hu et al., 2011):
(2n+1)H2 +nCO CnH2n+2 + nH2O (R.1)
2nH2 + nCO CnH2n + nH2O (R.2)
2nH2 + nCO CnH2n + 2O + (n-1)H2O (R.3)
CO + H2O CO2 + H2O (R.4)
The products of this process have been noted to follow the Anderson-Schulz-Flory (ASF) statistical hydrocarbon
distribution, where the DME-diesel fuel selectivity is 30% (Hu et al., 2012). Moreover, selection of the products is
also dependent on the reactor type and conditions employed for the process. There are two main types of FT
reactors including high temperature (HTFT) and low temperature (LTFT) systems, where the HTFT entails the use
of iron catalysts at temperatures of 613K to produce olefins and gasoline, and the LTFT utilizes iron/cobalt catalysts
at 503K to form diesel-based fuels and waxes (Hu et al., 2012). LTFT systems has been studied for the production
of DME, with the application of cobalt-based catalyst, which favours greater conversion, however in some instances
when the feed gas composition has a H2:CO stoichiometric ratio less than 2, the iron catalyst has been
recommended (Siedlecki et al., 2011). The reactor types typically used for FT processes include tubular reactors,
fixed bed, slurry and fluidised bed, most of which are also utilised in other DME production processes and will be
discussed later in the Chapter.
2.2.2 JFE Direct DME Synthesis Process
In 1989, one of the most prominent steel production plants in Japan commenced studies on the utilisation of coke,
a by-product from the plant, to produce DME. The company initial studies were lab-based producing very small
quantities, however over the years, the production capacity has since grown, with production levels of over 100

11
tonnes/day (Bourg, 2006). The figure below illustrates the present JFE DME production plant, with the attached
process pathway:

Figure 2.2. JFE Direct DME Production Plant with Production Flowchart (Bourg, 2006)

The JFE process was developed to produce DME in a single step, at conditions of 50 bars and 533K, with very
small catalyst particles containing three different active sites for the processes of methanol synthesis, dehydration
and water-shift gas reaction (Bourg, 2006). With this technique, natural gas is mixed with steam and oxygen to
produce syngas, where it then enters a slurry reactor to generate DME (Ohno et al., 2005). Within this system,
the reactions occur simultaneously, with the application of a bifunctional catalyst. The bifunctional catalyst contains
both the components required for methanol synthesis and dehydration, as well as the water shift gas reaction. One
of the most extensively used catalyst for the single stage process is the Cu-ZnO/-Al2O3 (Jalil et al., 2010). The
overall chemical reactions occurring within the reactor are expressed below (Azizi et al., 2014):
3CO + 3H2 CH3OCH3 + CO2 (R.5)
2CO + 4H2 CH3OCH3 + H2O (R.6)
The reaction 6 involves four stages of chemical interactions and are presented in the following chemical equations:
2H2 + CO CH3OH (R.7)
3H2 + CO2 CH3OH + H2O (R.8)
2CH3OH CH3OCH3 + H2O (R.9)
CO + H2O H2 + CO2 (R.10)
In the reactor system, reaction 9 utilizes the methanol produced by the reactions 7 and 8 (Jalil et al., 2010).
Additionally, the water formed in the reactions 8 and 9, employs the water-gas shift reaction mechanism to react
with CO, producing H2 and CO2 (Xie et al., 2015). The reactions and conversion rates of the chemical species with
the single stage process have been found to have close relationship with the H2: CO composition ratio. The figure
below depicts the equilibrium conversion of methanol and DME within the reactor at the selected operating
conditions. The figure shows the effect of the H2: CO ratio on the conversion of the compounds, where DME
experienced the greatest conversion at a stoichiometric ratio of 1, and methanol at a ratio of 2. In the figure, it is
also quite visible that the maximum equilibrium conversion of DME occurs at a greater degree than that of
methanol.

12
Figure 2.3. Equilibrium Conversion of Methanol and DME at 50 bars and 533K (Ohno et al., 2005)

Furthermore, the overall reaction within the system is exothermic, being dominated by methanol synthesis (Ogawa
et el., 2003). It is therefore, crucial to have temperature control as the processes and catalysts are highly
temperature-sensitive, and as such, mostly slurry type reactors are recommended. The JFE production processes
has been reported to produce 99.8% DME purity level fuel, which is usually mixed with LPG for utilization purposes
in Japan (Ohno et al.,2005). This technique has been studied by many other researchers and is therefore
considered one of the possible pathways for DME production in other parts of the world.
2.2.3 Two-Stage Production of DME
DME synthesis can be also be conducted in a two-stage process, including methanol formation followed by
methanol dehydration in individual reactors, with catalysts that are specific to each reaction (Laurence, 2009). This
production pathway entails the hydrogenation of the carbon oxides specifically CO and CO2, along with the reverse
water-shift gas reaction in a reactor, followed by a separate unit for the dehydration of the methanol. The following
equations describe the chemical reactions that occurs during the two-stage production of DME from syngas (Graaf
et al., 1988):
2H2 + CO↔ CH3OH Hydrogenation ∆H=-90.55kJ/mol (R.11)
3H2 + CO2 ↔ CH3OH + H2O Hydrogenation ∆H= -49.43kJ/mol (R.12)
CO2 + H2 ↔ CO + H2O Reverse Water Gas Shift Reaction ∆H= -41.12kJ/mol (R.13)
2CH3OH ↔ CH3OH + H2O Methanol Dehydration ∆H= -21.55kJ/mol (R.14)
In the first reactor, CO and CO2 are reacted with H2 to produce CH3OH, a mechanism that involves various stages
of formation as represented by Yang et al (2013). The possible reaction pathways for CO and CO 2 conversion to
methanol are expressed in the figure below:

13
Figure 2.4. Reaction Pathways for CO and CO2 to CH3OH (Yang et al., 2013)

According to researchers, methanol formation can be attributed to dual-site Langmuir Hinshelwood mechanism. In
a study conducted in 2010 (Yusup et al., 2010), it was reported that during catalytic methanol synthesis, CO and
CO2 becomes attached to an active site on the catalyst surface while H2 and H2O on another site, where the
hydrogenation of CO occurs. Yusup et al (2010) noted that the adsorption abilities of CO and H2 are dependent on
the partial pressures within the system. It was concluded in that study that methanol synthesis reactions are due
to a combined mechanism of adsorption, surface reactions and desorption (Yusup et al., 2010). Given the reactions
and desorption of methanol in the gas steam, a separation process is applied to remove the water and other
unreacted gaseous species. The methanol vapours are then transferred to a DME reactor, where catalytic
dehydration occurs. In the final stage of production, DME and water are formed, where the water and unreacted
methanol are then removed through a two-stage distillation process to produce a high-quality liquid biofuel (Farsi
et al., 2011).
2.2.4 Direct vs Indirect Production of DME
Studies have revealed that the production processes of DME, occurs at varying conditions, where the direct and
indirect differ greatly. The table below outlines the recommended operating conditions of the direct and indirect
production of DME:
Table 2.1. Operating Conditions for the Direct and Indirection Conversion of DME (Ohno et al., 2005)

Condition Direct Synthesis Indirect Synthesis


Methanol Formation Methanol Dehydration
Reaction Pressure (bars) 50 80-100 10-20
Reaction Temperature (K) 513-553 453-543 573-613

It is therefore, important to develop an understanding of the production pathways and the respective operating
conditions, in order to have favourable conversions and product compositions. Past studies have produced
debateable standings as against the two techniques, attesting to both the positives and negatives of each process.

14
Taupy et al (n.d) in a study related to the direct production, noted that the single stage yields greater conversion
rates producing larger amounts of DME. Jalil et al (2010) and Azizi et al (2014) confirm the results presented by
Taupy, however, Jalil et al also noted that the single stage pathway is highly exothermic since it combines three
exothermic reactions in a single reactor. Taupy et al (n.d) along with Laurence (2009), noted that the direct
production involves careful monitoring of the process to prevent issues such hot spots in the reactor and may
require costly capital investments and operating expenses. In a related study produced by Xie et al (2015), it was
indicated that the utilization of the bifunctional catalyst must be done with careful consideration to the catalyst type,
operating conditions and physical combination technique, factors that can greatly affect the performance of the
catalyst. Moreover, it was quite interesting that very few studies have been conducted on the reaction kinetics of
the direct production of DME, with more focus geared towards the two-staged process. Aziz et al (2014) in a study,
compared the direct and indirect production techniques and concluded that the direct stage is unfit for commercial-
level given the complexity in the separation system and unreacted gases in the flow stream which greatly affects
the conversion efficiency and product composition. In summary, the two-staged indirect process have been
reported to have greater flexibility and control with the chemical reactions, catalysts performance and deactivation
and operating conditions, since the processes are conducted in separate units (Wyoming, 2014).
2.3 Reactor Models
Given the operating conditions, and catalytic requirements of DME synthesis, the selection of the reactor type is a
critical consideration. Research has shown that three major types of reactors including the fixed bed, fluidised bed
and slurry reactors have been noted to produce acceptable DME conversion levels. Table 2.2 outlines some of the
major features of the three types of reactors:
Table 2.2. Properties of the Three Main Reactor Types used for DME Synthesis (Hu et al., 2011)

Properties Fixed-Bed Reactor Fluidised Bed Reactor Slurry Reactor


Temperature Control Poor Good Good
Heat Exchanger Surface (per 1000m3) 240m2 7-15m2 50m2
Catalyst Effects Small Large Intermediate
Back-Mixing Effects Low Intermediate High
Indicated Costs (GBP) 9023393.39 ---- 10529712.58

The table allows for a comparison of the major properties required for the selection of the most appropriate reactor
for optimal DME conversion. Literature has indicated that the abilities of the reactor to improve temperature control
and interactions between the catalyst and reactor gases contributes significantly to its application. In some cases,
the utilization of inert materials such as paraffin, with great thermal capacities allow for the capture and removal of
heat produced during the processes (Laurence, 2009). Additionally, adjustments to reactor structures can also be
applied for heat removal, with the shell and tube design as a highly favoured approach. The shell and tube
technique allows for the use of a coolant, mostly water, through the shell of the reactor, capturing the heat as it

15
moves along (Dadge et al., 2016). Figure 2.5 allows for a visualization of the basic structure of a slurry, fluidised
bed and fixed bed reactors:

Figure 2.5. Schematic of a Slurry, Fluidised Bed and Fixed Bed Reactors (Laurence, 2009)

Each reactor type has been strongly defended and reviewed in past studies, with the systems having advantages
and disadvantages attached to the operations. With the fixed-bed reactor, the catalyst is placed in tubes, where
the feed gas passes along, allowing for contact with the two components and enabling a catalytic reaction. In the
slurry reactor, however, the catalyst particles are dissolved in a high thermal capacity liquid, where the inlet gas is
then bubbled through to produce methanol and further DME. With respect to the fluidised bed reactors, these
systems can be constructed where the fluidised bed of catalyst is fixed or circulating, allowing the feed gas sufficient
contact time to react (Laurence, 2009).
Researchers have argued that the application of fixed bed reactors for DME synthesis is recommended, given its
simple design and ability for temperature control (Dagde et al., 2016). De Maria et al (2013) indicated that fixed
bed reactors are extensively used for industrial production of similar compounds, given the design to have attached
cooling tubes within the reactor, which can result in an isothermal system. On the other hand, Laurence (2009),
argued that slurry reactors have simpler design and reduced capital and operating costs as compared to the fixed
bed reactors. Contrary to this, Hu et al (2011), as shown in the table 2.2, stated that slurry reactors may have
greater costs attached to its application. Additionally, Laurence (2009), also indicated that slurry reactors allow for
better temperature control and prevents the formation of hot spots in the system. Wyoming (2014), however noted
that slurry reactors are quite large, requires great amounts of catalyst and may result in minimal mixing. Azizi et al
(2014) reported findings in agreement with Wyoming, where it was noted that due to the transfer of the gas to the
catalyst stream, followed by attachment on the catalyst particles, the mass transfer rates in the system may be
affected, resulting in reduced reaction rates. Added to this, is the case presented by Ogawa et al (2003), who
stated that conversion rates in the slurry system is highly dependent on the gas residence time, a factor that
requires careful consideration. With regards to the fluidised bed, these systems have been reported to be in the
testing phase with respect to application to methanol synthesis and dehydration processes. However, studies have
noted that the ability for improved temperature control, catalyst-gas mixing and heat removal in such systems may

16
be greater as compared to slurry reactors (Azizi et al., 2014). On the other hand, fluidised bed reactors have been
found to have issues relating to catalyst collisions and reactor wall effects, areas that may affect its application for
industrial DME production (Azizi et al., 2014).
The review of the literature has revealed that many studies recommend the use of fixed-bed reactors for DME
production, given its properties and structure, therefore the utilisation of the said system was selected for this study.
2.4 Catalysts Selection
In DME synthesis, the chemical reactions in the system can follow various pathways, whereby hydrogenation of
CO and CO2 may favour higher alcohols rather than selectivity of CH3OH and DME (Mayra et al., 2008). It is
therefore, crucial for the addition of catalysts to the reactor systems. Catalysts are compounds that increase the
occurrence of chemical reactions by reducing the energy requirements for such reactions. Catalysts, however may
only have selective influence on certain reactions, thereby increasing the selectivity of some products. Mayra et al
(2008) indicated that the structure, size and shape of catalyst particles can greatly affect its performance, lifetime
and stability.
In 1923, the very first methanol synthesis plant was constructed, where methanol was produced in a high pressure-
temperature system with the use of ZnO/Cr2O3 catalysts (Alam, 2013). However, as research progressed, low-
pressure, highly active and stable catalysts were developed for the methanol formation process. The Cu/ZnO
catalyst with the support of Al2O3 has been reported to have greater selectivity and stability, to produce methanol
(Rahman, 2012). In a study produced by Bakhtiary et al (2008), the possible composition of the Cu/ZnO/Al2O3
catalyst was presented, where Cu is reported to be 60-70%, ZnO as 20-30% and Al2O3 as 5-15%. Rahman (2012)
stated that the addition of the thermally stable Al2O3, support, to the catalyst improves its performance at various
operating conditions in the methanol synthesis process. In a paper written by Graaf et al (1988), it was related that
the Cu/ZnO/Al2O3 catalyst also enhances the activities of the water-gas shift reaction, given the acidic and porous
nature of the Al2O3 component. Xiao et al (2017) in a recent research, validates this notion, stating that the addition
of the Al2O3 support, not only catalyses the water-shift gas reaction, but also prevents the formation of unwanted
higher alcohols and increases the specific surface area of the catalysts. The figure below illustrates the effect of
various Al additions to the Cu/ZnO catalyst, which shows that the greater the Al, the higher the surface area (SBET)
and the smaller the pore diameter. The figure also depicts the NH3-TPD profiles for catalysts with varying amounts
of Al, whereby, the NH3 desorption peaks at 373-473 K and 573-773 K increases with increasing amounts of Al.
Xiao et al (2017) indicated that as this occurs, the level of acidity also increases, thereby enhancing the selectivity
for methanol formation.

17
Figure 2.6. Effects of the Catalyst Performance and Activities with Varying Amounts of Al2O3 (Xiao et al., 2017)

On the downside, Alam (2011), noted that the Cu/ZnO/Al2O3 catalyst is very selective to a CO-rich feed gas, with
reduced selectivity in the presence of CO2. Mayra et al (2008) confirms this statement, indicating that the increased
levels of CO2 may affect the activeness of the catalyst. However, the Cu/ZnO/Al2O3 catalyst is one the most
extensively used catalyst in the methanol production industry given its ability to produce high purity methanol at
low pressure-temperature conditions (Rahman, 2012 and Mayra et al, 2008). Bakhtiary et al (2008) noted that the
use of the Cu/ZnO/Al2O3 catalyst have resulted in 99.5% CO and CO2 conversion to CH3OH, a claim that has been
confirmed by Arthur (2010).
On the other hand, the methanol dehydration process favours an acidic component to enhance the reaction. The
-Al2O3 and zeolite ZM5 components have been studied by researchers, with varied arguments as to the better of
the two. The dehydration of methanol to produce DME is greatly affected by the acidity of the catalyst, whereby
unwanted product formation is a critical consideration. With this process, the acidity of the catalyst must be within
a specific range, given that a low acidity can result in poor conversion rates, whereas higher acidity can trigger the
formation of unwanted products (Wyoming, 2014). Xie et al (2015) argued that the zeolite catalyst is superior to
the Al2O3, due to improved stability and activity. However, other studies have related that the Al2O3 is an effective
catalyst in the conversion of methanol to DME, resulting in its wide-spread application in the DME production
(Dagde et al., 2016).
This research, therefore employs the use of the Cu/ZnO/Al2O3 and -Al2O3 catalysts for methanol synthesis and
dehydration, respectively, given that many of the kinetic models presented in the literature are based on
experimental validations conducted with these commercial catalysts.
2.5 Effects of Operating Conditions on DME Production
2.5.1 Methanol Synthesis: Effects of Temperature and Pressure
In a reactor system, there are various parameters which affects the conversion rate of the reactants, including
temperature and pressure. For methanol synthesis, the hydrogenation the carbon oxides are exothermic reactions,
whereby heat is released in the process. Liu et al (2003) noted that due to the exothermicity and reversible nature
of these reactions, lower temperature environments are favoured for product formation.

18
The figure below illustrates the conversion of CO at different temperatures for varying pressures. It can be seen
that the lower temperature of 525K produced higher conversion rates as opposed to 675K.

Figure 2.7. CO Conversions at Various Temperatures (Arthur, 2010)

Many researchers have presented studies indicating varying temperatures as the ideal range for methanol
synthesis. Yusup et al (2010) in a study, indicated that 510K is the selected temperature for maximum methanol
formation, however Mayra et al (2008), disagrees, and noted that 520K produces optimal conversion. Contrary to
this, Graaf et al (1988) and Villa et al (1985), produced kinetic models, which were validated experimentally and
argued that 473-543K is the recommended temperature range for methanol synthesis. Following the identification
of these kinetic models, specifically that presented by Graaf et al (1988), many researchers have since conducted
modelling, simulation and experimental studies in the field, which further verified and confirmed the findings (De
Maria et., 2013).
With respect to pressure effects, the opposite has been reported, whereby higher pressure produces greater
amounts of methanol. Traditionally, the production of methanol was conducted at very high pressures, however
with research and improved system designs, methanol is now synthesised at lower pressures of 50-100 bars
(Rahman, 2012). In a study conducted by Mayra et al (2008), the effect of pressure on the methanol formation was
investigated, producing results that confirms to the argument presented by Sharma (2010), who indicated that if
greater pressure is added to a system, the system will shift to the direction to reduce that pressure, in this case to
the right. Figure 2.8 illustrates the mole fractions of the reaction species of methanol synthesis, along with the
effect of pressure on the mole fraction of methanol:

Figure 2.8. Mole Fractions of Reactions Species along Reactor, along with the Effect of Pressure on Methanol
Formation (Mayra et al., 2008)

19
The figure shows that as pressure is increased, the mole fraction of methanol increases, therefore validating the
results presented by other studies. Alam (2011) argued that at 10 bars the conversion of syngas to methanol
achieves approximately 90%, and as this pressure grows, to about 80 bars, the conversion rate rises to almost
completion. Similar to that of temperature, Graaf et al (1988) and Villa et al (1985), also conducted studies with
varying pressures and methanol synthesis effects, where it was reported that a pressure range of 10-80 bars can
be applied.
Given the findings presented in literature, a sensitivity analysis of temperatures 473-563K and pressures 10-85
bars and the effects on methanol formation was one of the major focus of this research. The results were analysed
and discussed in Chapter 4.
2.5.2 Methanol Dehydration: Effects of Temperature and Pressure
The methanol dehydration thermodynamics is quite similar to that of the synthesis process, specifically with respect
to temperature. The dehydration of methanol to DME is an exothermic reaction, producing heat in the process.
Farsi et al (2011) stated that exothermic, reversible chemical reactions have an optimal temperature where
production is at its peak, any temperature increase above this point results in decreased product formation. Farsi
et al (2011), also indicated that for methanol dehydration, a lowered temperature would shift the equilibrium to the
right thereby increasing DME production. In a research conducted on isothermal and adiabatic reactors and DME
production (Farsi, 2015), it was reported that systems with enhanced temperature control resulted in greater
methanol conversion.
On the other hand, literature has shown that pressure has no effect on the equilibrium conversion of methanol to
DME. In the methanol dehydration, the chemical reactions have an equimolar stoichiometry, whereby the moles
of the reactants are equivalent to the moles of products, a state which allows the system to operate without any
impact from pressure changes (Farsi, 2015). Sharma (2010) confirms this statement, noting that the equilibrium of
the methanol dehydration to DME is not affected by any addition or reduction in pressure.
Based on the literature reviewed, the sensitivity analyses for the methanol dehydration in this study was therefore
done at temperatures 473-613K and pressure of 2.1 bars.
2.5.3 Isothermal Conditions for Methanol Synthesis and Dehydration
In typical reactors with adiabatic conditions, methanol synthesis and dehydration reactions would result in a rise in
temperature as the reactions occur along the reactor. According to Farsi et al (2011), in isothermal conditions,
temperature control is achieved through the use of cooling tubes, whereby heat is removed from the system
resulting in a constant temperature profile. The sensitivity of the methanol synthesis and dehydration processes to
temperature changes has resulted in numerous studies aimed towards the investigation and development of
isothermal systems. Alam (2011), reported that heat control in methanol synthesis is crucial, given the effects on
the equilibrium conversion. Alam (2011), as shown in figure 2.9, indicated that with isothermal systems, excess
heat removal can result in conversion efficiencies greater 90%, a rate that is highly favourable for production.

20
Figure 2.9. CO Conversion against Reaction Temperature in an Isothermal System (Alam, 2011)

In addition to this, the dehydration process has also been reported to favour isothermal conditions. Dagde et al
(2016) stated that heat removal during methanol conversion is critical to ensure equilibrium conversion to DME
and prevent catalyst deactivation. Given the findings of past studies, Farsi (2015) investigated the effects of DME
production in adiabatic and isothermal systems. Farsi (2015) reported that with isothermal systems, there is a 6.2%
greater methanol conversion as opposed to adiabatic systems. The figure below represents the findings presented
by Farsi (2015), illustrating greater DME production in isothermal conditions:

Figure 2.10. Mole Fractions of DME in Adiabatic and Isothermal Systems (Farsi, 2015)

Having established that an isothermal reactor is the ideal system for methanol and dehydration, the models
presented in this study were therefore, assumed to operate in isothermal conditions.
2.6 Kinetic Models for Methanol Synthesis
For the past few decades, numerous models have been reported in literature, each with kinetic equations that have
been validated against experimental data. These kinetic models, specifically for methanol synthesis are based
assumptions regarding the operating conditions and reaction mechanisms. The very first model for methanol
synthesis with the Cu/ZnO/Al2O3 catalyst was presented by Leonov et al in 1973. In this model, the effects of CO2
reactions were ignored and much focus was given to the hydrogenation of CO, noting it to be the key source of
carbon in the synthesis process (Bussche et al., 1996). In 1982, Klier et al., proposed a new model which included

21
the reaction of CO2, however the impact of the reaction on the kinetics were deemed minimal (Bussche et al.,
1996). Klier et al (Bussche et al., 1996) reported that the due to the high adsorption abilities of CO2, the reaction
rate of methanol synthesis and catalysts properties are greatly affected at varying PCO/PCO2 ratios. In this study, it
was further indicated by Klier et al, that the active, inactive and oxidised sites on the catalyst surface are directly
linked to the PCO/PCO2 ratio, where a Kredox term was added to reaction rate equation (Bussche et al., 1996). Villa et
al. (1985), also produced kinetic models whereby the hydrogenation of CO2 was excluded. In this model, low
pressure methanol synthesis was conducted using the Cu/Zn/Al2O3 catalyst, where the hydrogenation of CO and
the reverse water-gas shift reaction were noted to be the main reaction pathways in the system. Villa et al (1985),
reported that the due to the adsorption capacity of the CO2, the additional reverse water gas shift reaction would
occur and should therefore be accounted for in the kinetics. Contrary to these studies, Graaf et al (1988)
investigated the effects of CO, CO2 and the water-gas shift reactions during low-pressure methanol synthesis over
the Cu/ZnO/Al2O3 catalyst. Graaf et al (1988) reported that due to the occurrence of the water-gas shift reaction,
the reaction mechanism for methanol synthesis includes both the hydrogenation of CO and CO 2. In this study,
Graaf et al (1988) examined the specific mechanisms for the three reactions, specifically hydrogenation of the
carbon oxides and the water-gas shift reaction, where the rate controlling steps were used to derive the kinetic
equations. Graaf et al (1988) produced 48 kinetic rate models for the three main reactions and indicated that the
water-gas shift reaction is much slower than the methanol formation reactions, where much of the water in the
product stream is as a result of the CO2 hydrogenation. The overall reactions, with the elementary steps and driving
force as described by Graaf et al (1988) are presented below:
(A) CO + 2H2 = CH3OH (R.15)
(B) CO2 + H2 = CO + H2O (R.16)
(C) CO2 + 3H2 = CH3OH + H2O (R.17)

Figure 2.11. Reactions and Driving Force for Methanol Synthesis (Graaf et al., 1988)

Using experimental studies, Graaf et al (1988), proved that the formation of methanol is due to both CO and CO2
in a comprehensive examination, where statistical analysis was applied to derive the three main kinetic equations
for the model.

22
The methanol synthesis models discussed above are presented in table 2.3, where the operating conditions and
kinetic rate equations are expressed:
Table 2.3. Initial Kinetic Models for Methanol Synthesis

As research progressed, many other methanol synthesis models were developed. In 1995, Askgaard et al.,
produced a model based on gas-phase thermodynamics and surface science. In this study, 100 experiments were
conducted, where the rate limiting step in the kinetics was found to be the hydrogenation of H2COO to methoxide
and oxide (Askgaard et al., 1995). Askgaard et al (1995) also included the effects of the water-gas shift reaction,
where it was noted that the water-gas shift reaction is in competition with the methanol synthesis reaction on the
surface of the catalyst. According to Askgaard et al (1995), H2 and formate molecules were the prevalent species
in the system, where formate was found to be a highly stable and active component. It was reported that at
temperatures below 500 K, the highly active formate begins to bombard the free sites on the catalyst, resulting in
interference of methanol formation (Askgaard et al., 1995). Askgaard et al (1995) indicated that the results of this
study were proven satisfactory with experimental data. In the following year, Bussche and other colleagues,
produced a steady-state kinetic model for methanol synthesis over the Cu/ZnO/Al2O3 catalyst, which included the
sensitivity of the kinetics to pressure, temperature and CO: CO2 ratio (Bussche et al., 1996). In this study, Bussche
et al (1996), indicated that methanol formation is mainly due to the hydrogenation of CO 2, where the water-gas
shift reaction also plays an important role in the kinetics. Bussche et al (1996) concluded that the kinetic rates
produced in this study were based on the overall reactions being examined through common surface oxygen
intermediates. Additional to this, is another study proposed by Kubota et al (2001), where Cu/ZnO catalyst was
applied in the methanol synthesis reactions. Kubota et al. (2001) confirms to the findings of Bussche et al, reporting
that methanol formation is as a result of the hydrogenation of CO2. It was indicated in this paper that the methanol

23
formation is mainly through the intermediates of formate and methoxy species, whereby the reaction between the
formate and H2 on the Cu/Zn catalyst was the rate determining step in the system (Kubota et al., 2011). It was
noted that the kinetic models produced by Kubota et al, were tested experimentally and indicated a 10% difference
with industrial-level production data (Kubota et al., 2001). One of the most recent models proposed for methanol
synthesis was reported by Lim et al. in 2009, where the production was catalysed by Cu/ZnO/Al2O3/ZrO2. In this
study, the effects of CO, CO2 and water-gas shift reactions were investigated in a comprehensive setting. Lim et
al (2009) produced 48 reaction mechanisms from varying combinations of rate determining steps and found three
main rate determining pathways for the reactions of CO, CO2 and water-gas shift. The surface reaction of the
methoxy compound, hydrogenation of a formate intermediate HCO2 and formation of a formate intermediate were
reported as rate determining steps for the CO, CO2 and water-gas shift reactions, respectively (Lim et al., 2009).
Lim et al (2009) also noted that the hydrogenation of CO2 occurs at a slower rate as compared to CO and that CO2
levels decreases as the water-gas-shift reaction occurs.
The models that were later presented in literature are shown in the table 2.4, where the operating conditions and
kinetic rate equations are expressed:
Table 2.4. Additional Kinetic Models for Methanol Synthesis

The model produced by Graaf et al in 1988 has been found to be most extensively used for methanol synthesis,
given the experimental validation, which indicates significant agreement with lab-scale and industrial-level
production data. The model has also been studied by many researchers including Alam (2011), De Maria et al
(2013) and Yusup et al (2010), whom have all reported that the model confirms to experimental results. Many
studies have also validated the thermodynamic equilibrium constants, adsorption equilibrium constants, reaction

24
rate constants and reaction rate equations reported by Graaf et al. Therefore, in this study, the kinetics reported
by Graaf et al in 1988 was selected to model the hydrogenation of CO and CO2 and the water -gas shift reaction
in the methanol synthesis process.
2.7 Kinetic Models for Methanol Dehydration
Much like the methanol synthesis, many kinetic models have also been proposed for the dehydration of methanol
to DME. One of the very early models, presented by Bercic et al. (1993), was based on a one-dimensional,
heterogeneous plug flow reactor with the utilization of 0.003 m -Al2O3 catalyst particles to dehydrate methanol.
According to Bercic et al. (1993), the intra-particle mass transport was the rate controlling mechanism for this
reaction. Bercic et al. (1993) produced a model for the rate per catalyst particle using concentrations of the reaction
species and other reaction parameters. It was also reported that Bercic et al. altered the rate equation by
expressing the concentrations in the form of mole fractions, where he applied apparent coefficients in the reaction
equation. Further to this, Sierra et al (2013) proposed a methanol dehydration mechanism with the use of the -
Al2O3 catalyst, where the kinetics were based on mass conservation in an isothermal, plug-flow system. In this
study, 13 models were presented, where the best model was found to be that which included the adsorption
capabilities of methanol, water and DME to the active sites on the catalyst (Sierra et al., 2013). Sierra et al (2013)
also stated that the methanol dehydration reaction is elementary and that the effect of water in the reaction
mechanism is critical to the kinetics. Additionally, in 2013, Zhang et al also investigated the dehydration of methanol
to DME using the -Al2O3 catalyst. Zhang et al (2013) presented a different kinetic model as compared to Sierra et
al, whereby the Langmuir-Hinshelwood surface-controlled reaction was expressed in the kinetics. Zhang et al
(2013) reported that this mechanism along with dissociative adsorption confirmed with experimental data and is
therefore a good representation for methanol dehydration. In this study, Zhang et al (2013), also noted that the
temperature and molar flow rates greatly affect the dehydration of methanol as opposed to pressure which was
found to have minimal impact on the system. Moreover, in 2013, another study proposed by Ziyang et al., was
presented for the methanol dehydration process. In this study, the -Al2O3 catalyst was also used to model a plug
flow reactor for the dehydration process. Ziyang et al (2013), indicated that the model presented was based on the
fugacity of the compounds and other reaction parameters, where it was found to be in satisfactory agreement with
experimental data.
The models for the methanol dehydration process that were reviewed are displayed in the table below, where the
operating conditions and rate equations are presented:

25
Table 2.5. Kinetics Models for Methanol Dehydration Presented in Literature

Given the findings and validations of the models presented, the kinetics proposed by Bercic et al in 1993 was
selected for this study. This model has been studied and tested by many researchers including Farsi et al (2015)
and Dagde et al (2016), where it was proven to produce valid and experimentally fit results for the dehydration of
methanol to DME.
2.8 Conclusion
The production of DME has been investigated by many researchers, whom have proposed various techniques,
models and conditions for ideal equilibrium production. The direct and indirect methods including the FT, JFE and
two-stage methanol synthesis-dehydration techniques are currently employed in various countries for the
production of DME. The selection of reactor type and catalyst combinations has also been critical considerations
for the production processes, given the sensitivity, costs, and conversion factors related to DME synthesis. The
two-stage production process along with an isothermal-fixed bed reactor was selected for this study with the
application of the Cu/ZnO/Al2O3 and -Al2O3 catalysts. The selection of this technique was based on the literatures
presented and reviewed in this chapter. Moreover, the kinetics for the modelling of methanol synthesis and
dehydration have seen much research from the early 1990’s. The many models varied greatly with respect to
catalyst application, operating conditions and reaction mechanisms. Based on experimental and literature
validation, the models proposed by Graaf et al (1988) and Bercic et al (1993) were selected for this study. The
conditions, rate equations and reaction pathways were presented and reviewed in this chapter.
Given the literature presented, the methodology and selected theories for this study is discussed in the following
chapter.

26
Chapter 3: Methodology
This chapter describes the methodologies that were applied to achieve the objectives set out by the study. Several
techniques including literature reviews, mathematical modelling and sensitivity analysis tools were utilised to
accomplish the goals of this research.
3.1 Introduction
The completion of this study was attained with the application of qualitative and quantitative research strategies.
The preliminary approach was focused on the identification and examination of the present and future demand
level of renewable alternatives for diesel fuel in India. Additional to this, an understanding of the kinetics, related
theories and past findings of methanol synthesis and dehydration models were also crucial to achieving the
objectives of this research. In conducting the qualitative research, content analysis was utilised, which included
the study of related journals, books, scientific notes, conference papers and pilot projects. The analysis of past
research allowed for the identification of trends, relationships and possible gaps in the related fields. Much of this
data has been reviewed in the previous chapter, however the kinetics and reaction mechanisms selected for this
study are discussed in this chapter with specific emphasis on the application to biomass-derived syngas from India.
The methodology is therefore divided in various sections including:
➢ Methanol Synthesis- Process Flow and Kinetic Model
➢ Methanol Dehydration- Process Flow and Kinetic Model
➢ Reactor Specifications
➢ Sensitivity Analysis- Temperature, Pressure and Feed-Gas Composition
The modelling and simulation analysis for this study was conducted with the use of Microsoft Excel, Aspen Plus
V8.8 and NASA CEA Equilibrium Programmes. The application of the programmes with the respective kinetics
and mathematical equations are discussed further, in the chapter.
3.2 Methanol Synthesis
3.2.1 Process Pathway
The proposed process pathway for the synthesis of methanol was constructed using Aspen plus V8.8 software.
The R-Gibbs property model for an ideal gas reactor was applied to produce a model representative of that
presented in the literature and that which can possibly be applied for biomass-based methanol production in India.
The process flowsheet allows with a simplified visual representation of the pathway is presented in the following
chapter.
3.2.2 Feed-Gas Composition
The inlet feed gas composition for this study was critical to the reaction kinetics and mechanisms. This research
was aimed towards the utilisation of India’s indigenous resource, agricultural wastes to produce syngas via
anaerobic digestion and partial oxidation, which is then synthesised to methanol and further dehydrated to DME.
The syngas used in the analysis has been selected based on partially-oxidised methane gas compositions. The
composition of syngas applied in this work is shown in the table below:

27
Table 3.1. Syngas Composition used in the Study

Composition Mole Percent


CO 10.4
CO2 5.6
H2 83.5
H2O 0.5

3.2.3 Kinetics of Methanol Synthesis


In this study, the kinetics for the methanol synthesis was sourced from Graaf et al (1988). The model was computed
based on various assumptions that have been proposed including:
➢ Single-pass shell and tube, fixed-bed catalytic reactor was selected.
➢ Plug flow model was used, where the gases were assumed to flow through the reactor with no radial velocity
gradient.
➢ Isothermal conditions were applied.
➢ Axial and radial dispersion were ignored.
➢ Ideal gas phase was applied.
➢ Steady-state conditions were considered.
➢ Effectiveness factor was assumed to be one.
➢ Intra-particular diffusion was not included.
The chemical reactions that were used to derive the kinetic models presented in this study are expressed below:
CO + 2H2 CH3OH (R.18)
CO2 + 3H2 CH3OH + H2O (R.19)
CO2 + H2 CO + H2O (R.20)
The kinetic models for the hydrogenation of CO and CO2 and reverse water-gas shift reaction, based on the
findings of Graaf et al (1988):

 
 3 f CH3OH 
k1 K CO  f CO f H 2 
 
2
1 
 f H 2 2 K P1 
rCO   
   K  
 
1  K CO fCO  K CO2 f CO2  f H 2 2     fH O  
1 H 2O
   K 12  2   (3.1)
   H2  

 3 f CH 3OH f H 2O 
k2 K CO2  f CO2 f H 2 2  
  
f H 2 3/2 K P 3 
rCO 2 
  K H 2O  
 
1  K CO f CO  K CO2 f CO2  f H 2 1/2  
 K H 1/2  H 2 0 
 f (3.2)
  2  

28
 f H O fCO 
k3 K CO2  fCO2 f H 2  2 
rH 2O   KP2 
(3.3)
  K H2 0  
 
1  K CO f CO  K CO2 f CO2  f H 2 1/2  
 KH 1/2  f H 2O 
  2  

For each of the reaction rates, the reaction rate constants were calculated using the equations presented by
Graaf et al (1988):
 109900  (3.4)
K1  2.69  107  exp  
 RT 
 123400 
K 2  7.31108  exp   (3.5)
 RT 
 65200 
K 3  4.36 102  exp   (3.6)
 RT 

The results for the reaction rate constants used to derive the rates of the chemical reactions are shown in the
table 3.2, where T=483K and R=8.314 Jmol-1K-1:
Table 3.2. Reaction Rate Constants for Methanol Synthesis

Reaction Rate Constants k


k1 3.505×10-05
k2 3.877×10-05
k3 3.303×10-05

Additional to this, the adsorption equilibrium constants required in the rate equation, were also derived using the
equations presented by Graaf et al (1988):

 58100 
K CO  7.99 107  exp  
 RT  (3.7)

 67400 
KCO 2  1.02 107  exp   (3.8)
 RT 
K H 2O  104500 
1  4.13 1011  exp  
KH 2 2  RT  (3.9)

By the application of the above equations, the adsorption equilibrium constants were computed and are
presented in table 3.3, where T=483K and R=8.314Jmol-1K-1:

29
Table 3.3. Adsorption Equilibrium Constants for Methanol Synthesis

Adsorption Equilibrium Constant, K


KCO
1.53
KCO2
1.98
(KH2O/H2)1/2
8.26

The reaction equilibrium constants for the methanol synthesis were also sourced by Graaf et al (1988), where the
following equations were employed:
(3.10)
 12.621  P in bars 
5139
Log10 Kp1 
T

2073 (3.11)
Log10 Kp2   2.029  P in bars 
T

Kp3  Kp1  Kp2 (3.12)

The following results were derived for the reaction equilibrium constants, where T=483K and were applied in the
determination of the reaction rates:
Table 3.4. Reaction Equilibrium Constants for Methanol Synthesis

Reaction Equilibrium Constant Kp


Kp1
1.0×10-02
Kp2
5.45×10-03
Kp3
5.69×10-05

In the derivation of the reaction rates, the fugacity (ƒ) were assumed to be partial pressures of the respective
compounds, given that the gases were assumed to be ideal. The mole fractions and total pressure of the reactor
system were used to determine the partial pressures of the species along the reactor length, the relationships are
described in the table below:
Table 3.5. Mole Fraction and Partial Pressure Relationships

Inlet Mole Fraction Consecutive Mole Fractions Partial Pressures

Υi, inlet =
𝑀𝑜𝑙𝑒 % 𝐹𝑖
Υi = 𝐹𝑇 Pi = Υi × PT
100

Given the partial pressures and mole fractions were known, the rates of the reaction at the first iteration were
calculated based on equations 18, 19 and 20. In order to compute the rates of reaction along the reactor length,

30
the rate of change in flow rates and the respective flow rates were calculated. The rate of change in the flow of
methanol was firstly determined as shown below:
dFCH3OH = [(FCoin - FCOeqm) + (FCO2in- FCO2eqm)] / no. of iterations (3.13)
The FCOeqm and FCO2eqm were derived using the NASA CEA Programme where the syngas composition and
operating conditions (temp and pressure) were entered to produce the probable mole fractions of the reactants
and products after equilibrium conversion.
The NASA CEA equilibrium conversion of CO and CO2 were computed as follows:
Table 3.6. NASA CEA Equilibrium Conversion of CO and CO2

CEA Conversion (%) FCO,eqm FCO2,eqm

Conv. (%) =
𝛶𝑖,𝑖𝑛𝑙𝑒𝑡−𝛶𝐶𝐸𝐴,𝑒𝑞𝑚
× 100 FCO,eqm= Conv. % × FCO,inlet FCO2,eqm= Conv. % × FCO2,inlet
𝛶𝑖,𝑖𝑛𝑙𝑒𝑡

The rate of change and the flow rates of the other reaction species along the length of the reactor were determined
using the following relationships:
Table 3.7. Rates of Change in Flow Rates and Respective Flow Rates of Reaction Species along Reactor

Species Rate of Change in Flow Flow Rate along Reactor End of Iteration Flow Rate
CO dFCO= (-r1 + r3) × dW FCO= FCO(previous iteration) + dFCO FCO,enditer= FCO + dFCO
CO2 dFCO2= (-r2 -r3) × dW FCO2 = FCO2(previous iteration) + dFCO2 FCO2,enditer= FCO2 + dFCO2
H2 dFH2 = (-2r1-3r2-r3) × dW FH2 = FH2(previous iteration) + dFH2 FH2,enditer= FH2 + dFH2
H2O dFH2O = (r3 + r2) × dW FH2O = FH2O(previous iteration) + dFH2O FH2O,enditer= FH2O + dFH2O

In order to determine the rates of change in the flow rates of the species, the change in the weight of the catalysts
and the total weight for the synthesis of methanol were determined as follows:
Table 3.8. Rate of Change in Weight of Catalyst

Rate of Change of Catalyst Weight (kg) Catalyst Weight (kg)

dW =
𝑑𝐹𝐶𝐻3𝑂𝐻 W= dW + W
𝑟1+𝑟2

It should be noted that the rates of change in the flow rates of the respective compounds were formulated based
on the chemical reactions and relations expressed in chemical equations 18, 19 and 20 discussed earlier in the
chapter.
The final aspect of the methanol synthesis model was focused on the determination of the conversion rates of CO
and CO2 as predicted in the model. This analysis allows for an understanding of the efficiency of the production
process as proposed by the kinetic model, aiding in its applicability for commercial level. The conversion rates in
% were calculated as shown in the table below:

31
Table 3.9. Conversion Rates of CO and CO2 as Predicted by the Model

Model: CO Conversion % Model: CO2 Conversion %


𝐹𝑐𝑜,𝑖𝑛−𝐹𝑐𝑜,𝑒𝑛𝑑𝑖𝑡𝑒𝑟 𝐹𝑐𝑜2,𝑖𝑛−𝐹𝑐𝑜2,𝑒𝑛𝑑𝑖𝑡𝑒𝑟
CO =( 𝐹𝑐𝑜,𝑖𝑛
)×100 CO2=( 𝐹𝑐𝑜2,𝑖𝑛
)×100

The model predicted CO and CO2 conversion rates were compared to that produced by the NASA CEA
equilibrium software and the results are presented in the following chapter.
3.3 Methanol Dehydration
3.3.1 Process Flow
A process flowsheet for the dehydration of methanol to DME was developed using the Aspen Plus V8.8 software.
The model was designed based on industrial-level productions that were described in the literature (Farsi et
al.,2015). Like the methanol synthesis process, the model was based on the R-Gibbs property model for an ideal
state, which displays the pathway for methanol dehydration to DME. The process sheet is presented in Chapter 4.
3.3.2 Feed-Gas Composition
Similar to the methanol synthesis process, the feed-gas composition was also important for the reaction of
methanol dehydration. The simulation of the ideal methanol dehydration model was based on an inlet gas stream
with the molar composition of CH3OH being 1, where water and other unreacted gases from the methanol synthesis
process were extracted (Refer to process flow sheet in Chapter 4).
3.3.3 Kinetics of Methanol Dehydration
The kinetics and reaction pathways employed for the dehydration of methanol to DME, were sourced from Bercic
et al (1993). In the application of this model, various assumptions were derived and are listed below:
➢ Single-pass catalytic shell and tube fixed-bed reactor was chosen for the model.
➢ Plug flow was applied.
➢ Isothermal conditions were assumed.
➢ Axial and radial distribution were excluded.
➢ Ideal gas phase was applied to the model.
➢ Steady-state conditions were assumed.
➢ Catalyst effectiveness was taken as one.
The chemical equation considered in the simulation of the methanol dehydration model is shown below:
2CH3OH CH3OCH3 + H2O (R.21)
The equation shows that two moles of methanol produces one mole of DME and one mole of water. The kinetic
model for this chemical reaction that was selected is displayed below:

 yCH3OCH3 yH 2O 
kK 2CH3OH  yCH3OH 2  
 K eqm
r  

 
(3.14)
 
0.5 4
1  2 KCH3OH yCH3OH  K H 2O yH 20

32
The kinetic constants expressed in the rate equation were computed using equations that were also proposed by
Bercic et al (1993) and are presented in table 3.10
Table 3.10. Kinetic Constants used in the Methanol Dehydration Model

Kinetic Parameter Equation


Kinetic Rate Constant  10800 
k  6.60 108 exp  
 T 
Equilibrium Constant
ln  Keq   0.86log T 
3138
 1.33 103 T  T 2  3.5 1010 T
T
Adsorption Equilibrium  830 
K CH3OH  0.72  102 exp  
Constant  T 

 1130 
K H 2O  0.45 102 exp  
 T 

With the application of the above equations, the values of the kinetic parameters used in the methanol dehydration
rate model were determined, where T=573K and are displayed in table 3.11:
Table 3.11. Kinetic Parameters used in the Selected Model

Kinetic Parameters
1.19
k
96.7
Keq
KCH3OH 3.06×10-02
KH2O 3.23×10-02

To determine the rate of the reaction, the mole fractions of the reaction species were calculated based on the
equations depicted below:
Table 3.12. Mole Fraction and Partial Pressure Relationships

Inlet Mole Fraction Consecutive Mole Fractions Partial Pressures

Υi, inlet =
𝑀𝑜𝑙𝑒 % 𝐹𝑖
Υi = 𝐹𝑇 Pi = Υi × PT
100

In order to derive the consecutive mole fractions along the length of the reactor, the rate of change of molar flow
rates were computed as shown below:

33
Table 3.13. Rate of Change in Molar Flow Rates and the Respective Flow Rates of Reaction Species

Species Rate of Change in Flow Flow Rates along Reactor End of Iteration Flow Rates
Rates
CH3OH dFCH3OH= -2× r ×dW FCH3OH= FCH3OH, previous iter. + dFCH3OH FCH3OH,enditer= FCH3OH + dFCH3OH
CH3OCH3 dFCH3OCH3= (FCH3OCH3eqm × 0.5) FCH3OCH3= FCH3OCH3, previous iter. + FCH3OCH3,enditer= FCH3OCH3 +
/no. of iterations dFCH3OCH3 dFCH3OCH3
H2O dFH2O = (r × dW) × 0.5 FH2O= FH2O, previous iter. + dFH2O FH2O,enditer= FH2O + dFH2O

For the formation of DME, the equilibrium conversion of methanol at the given operating conditions were deduced
using the NASA CEA equilibrium programme. The equilibrium conversion rate and molar flow rates were found
using the equations in table 3.14:
Table 3.14. NASA CEA Equilibrium Conversion of Methanol

CEA Conversion (%) Equilibrium Flow Rate of CH3OH

Conversion (%) =
𝛶𝑖,𝑖𝑛𝑙𝑒𝑡−𝛶𝐶𝐸𝐴,𝑒𝑞𝑚
× 100 FCH3OHeqm= Conv. % × FCH3OH,inlet
𝛶𝑖,𝑖𝑛𝑙𝑒𝑡

Given the determination of the rate of change of DME formed along the reactor, the weight of catalysts employed
for this change was calculated using the following relationships:
Table 3.15. Rate of Change in Weight of Catalyst and Total Catalyst Weight along the Reactor Length

Rate of Change in Weight of Catalyst (kg) Total Weight of Catalyst (kg)

dW=
𝑑𝐹𝐶𝐻3𝑂𝐶𝐻3 W= dW + W
𝑟

As indicated with the methanol synthesis process, the rates of change in the molar flow rates of the reaction species
of the methanol dehydration model were also based on the chemical reaction in the system, specifically R.21.
Given the establishment of the flow rates and respective changes along the reactor, the conversion rate of
methanol to DME were derived. The methodology applied in this aspect is expressed in the equation below:
𝑭𝑪𝑯𝟑𝑶𝑯,𝒊𝒏−𝑭𝑪𝑯𝟑𝑶𝑯,𝒆𝒏𝒅𝒊𝒕𝒆𝒓
CH3OH Conversion % = ( 𝑭𝑪𝑯𝟑𝑶𝑯,𝒊𝒏
) ×100 (3.15)

The conversion rate of methanol was compared to the predictions presented by the NASA CEA equilibrium
programme and is discussed further in the chapter 4.
3.4 Reactor Specifications and Flow Properties
To compute the reactors specifications, the respective catalyst densities used in the methanol synthesis and
dehydration process were required. Table 3.16 displays the densities for the Cu/ZnO/Al2O3 and -Al2O3 catalysts
that were applied in this simulation study:

34
Table 3.16. Catalyst Densities of Methanol Synthesis and Dehydration Catalysts

Catalyst Type Density (kgm-3) Reference


Cu/ZnO/Al2O3 1770 De Maria et al. (2013)
-Al2O3 1130 Park et al. (2000)

The catalyst densities were selected based on the experimental data that has been provided in the literature with
respect to its application in similar processes.
Additional to the catalyst properties, the reactor diameter was also crucial in the computation of the reactor length.
For plug flow models, it is recommended that for ideal flow properties, the diameter of the reactor should be at
least ten times the diameter of the catalyst particles used in the system. A scenario as such, allows for greater
activity and reduced impacts of the greater velocities along the rector walls (Wilkinson, 2014). For the methanol
synthesis and dehydration systems, a catalyst particle of 0.003 m was selected based on findings presented in
literature (Bercic at al.,1993 and De Maria et al.,2013). Additionally, in such plug flow systems, it is also noted that
a rule of thumb applies for the length/diameter relationship, whereby the L/D should be approximately equal to
5/10. This range has been reported in literature for ideal conditions reducing the occurrence of pressure drops and
temperature gradients within the reactor. Given the materials presented in the literature, a L/D relationship for the
methanol synthesis and dehydration models were computed for the selected range of 5-10. The simulation of this
aspect was done with the use of Microsoft Excel and the results are presented in Chapter 4.
The reactors lengths for the methanol synthesis and dehydration processes were then calculated using the
equation below:
𝑾
Length of Reactor, L= (3.16)
𝑨×𝝆

Where W= weight of catalysts (kg); A= Area of Reactor (m2) and ρ is the density of the catalyst (kgm-3).
And A= πr2; where r=radius of reactor (m).
Additional to the reactor length, the volumetric flow rate and the velocity of the gas stream along the reactor was
determined. The ideal gas law was applied for this simulation as shown below:
PV= FRT; where V= FRT/P
Therefore, the Volumetric Flow Rate, Q was found using Q (m3s-1) = FRT/P. (3.17)
Given the volumetric flow rate and the area of the reactor, the velocity of the gas was found using the following
equation:
Velocity, v (ms-1) = Q × A (3.18)
The simulation of the volumetric flow rates and velocity of gas stream were conducted for both processes and the
results are presented in the next chapter.
3.5 Sensitivity Analyses: Effects of Temperature, Pressure and Feed-Gas Composition
The final aspect of this research was aimed towards conducting sensitivity analyses of the reactor models for the
methanol synthesis and dehydration processes. Sensitivity analysis of models produces crucial information

35
regarding the inputs variables in the system. This technique allows for an understanding of the output parameters
to uncertainties, resulting in the development of reliable and accurate reactor models.
Literature has indicated that the reaction systems for methanol synthesis varies with temperature and pressure. It
has been reported in many studies that due to the exothermicity of the reactions, the effects of high temperatures
can lead to unwanted product formation and unfavourable product composition. For this study, a temperature range
of 473-563K were tested for the methanol synthesis model. Similar to temperature, the effects of pressure have
also been indicated in previous studies, and as such a pressure range of 10-85 bars has been applied to the model.
The methanol produced during the synthesis process must be kept in the vapour phase in order to move along to
the dehydration reactor; therefore, the effects of low temperatures and high pressures on the vapour fraction of
methanol were determined using the Aspen plus V8.8 software. This analysis produced an insight of the
temperature-pressure range that would allow for the methanol produced to remain in the vapour phase.
The methanol dehydration model was also tested against varying temperatures, specifically 473-613K, a range
that was selected based on the literature reviewed. Pressure effects on the methanol dehydration model were
excluded in this study, given the equimolar state of the reaction and the independence of the reaction rate to partial
pressures.
The reactor models were also tested with varying compositions of the feed-gas in order to determine the best
possible inlet gas composition for the processes. For methanol synthesis, the feed-gas composition was varied
based on possible syngas production techniques, producing different stoichiometric ratios of H2:CO. The
composition of the inlet gas for the dehydration process were also tested with the inclusion of water and unreacted
gases. The effect on the composition of the feed-gas on the equilibrium conversion of CO, CO2 and methanol were
tested using the NASA CEA equilibrium programme. Further effects relating to catalyst weight, reactor
specifications and gas flow were computed using Microsoft Excel.
Moreover, the reactor models were simulated at industrial levels, where the molar flow rates were adjusted so as
to deduce the reactor details and catalyst weight required to scale-up the production processes.
3.6 Conclusion
The synthesis and dehydration of methanol to produce DME were modelled using the studies presented by Graaf
et al (1988) and Bercic et al. (1993) respectively. The kinetic models, reaction constants and parameters employed
in this study were presented in this chapter. The application of the specific models to deduce the flow rates, catalyst
weights and length of the reactors were discussed. Moreover, the approaches for the sensitivity of the reactor
models to the effects of temperature, pressure and feed-gas composition were also reviewed. The methodologies
discussed in this chapter has been applied for the methanol synthesis and dehydration processes and the results
are displayed and analysed in the following chapter.

36
Chapter 4: Results & Discussion
This chapter includes the presentation, analysis and discussion of the results that were derived during this
research. The applied methodologies in mathematical modelling, kinetics and simulation works resulted in the
following outcomes, which have been expressed in two sections; methanol synthesis and dehydration.
4.1 Methanol Synthesis
4.1.1 Process Flow of Methanol Synthesis
The kinetic model proposed by this study was an isothermal, steady-state, plug-flow system for a shell and tube
catalytic fixed bed reactor. The production of methanol has been reported to be limited by thermodynamics which
affects equilibrium conversion resulting in low conversion rates of CO and CO2 (Mayra et al., 2008). Therefore, in
this system, an attached recycling unit was proposed. In this system, as the inlet gas enters, a mixer is employed
to combine the recycled unreacted gases with the inlet feed, after which a compressor is used to increase the
pressure of the stream. The gas stream is the passed through a heater, where the desired temperature is
established. The gas then enters the recommended reactor which is divided into shell and tubes, in this system
3000 tubes were modelled. The catalyst, Cu/ZnO/Al2O3 is packed in the tubes in the reactor, where the shell side
is used as a passageway for a cooling liquid. The movement of the inlet gas is directed through the tubes in a
manner that allows for a well distributed flow stream. As the reactions occur, the outlet gas leaves the reactor and
is passed through a membrane separator which removes the unreacted gases and unwanted products, resulting
in a pure methanol final product. The unreacted gases are transferred to the mixer and the production process
continues. The proposed process flow is expressed in the figure below, where the relevant components are
presented:

Figure 4.1. Hypothetical Process Flowsheet of the Methanol Synthesis Process

4.1.2 Methanol Synthesis Model


This section presents the results for the ideal model that was developed using the kinetics presented by Graaf et
al (1988). The kinetics were applied in Microsoft Excel to produce models for the synthesis of methanol from the
hydrogenation of CO and CO2 and the reverse water-shift gas reaction. The ideal model for the synthesis as
prescribed by the kinetics was selected based on numerous simulations of models with varying characteristics and
input parameters. The main findings of the recommended model are presented in table 4.1, where reactor, catalyst
and flow stream properties are shown:

37
Table 4.1. Main Characteristics of the Selected Methanol Synthesis Model

Property Unit Value


Inlet Temperature K 483
Pressure Bar 85
Number of Iterations ---- 500
Number of Tubes ---- 3000
Reactor Length m 0.16
Reactor Diameter m 0.019
Length/Diameter of Tube ---- 8.21
Catalyst Particle Diameter m 0.003
Catalyst Density kgm-3 1770
Catalyst Weight kg 0.08
Total Inlet Molar Flow Rate mols-1 0.64
Velocity of Gas Stream ms-1 3.55×10-04
Volumetric Flow Rate m3s-1 1.01×10-07

The system was modelled at a lab-scale, using a single-pass shell and tube reactor design containing 3000 tubes.
The total flow rate was assumed to be 0.64 mols-1, passing through the system, resulting in a 0.0002 mols-1 molar
flow in individual tubes. Using the methodologies discussed in the previous chapter, the reactor length, catalyst
weight, velocity and volumetric flow rate of the system were computed and is presented in table 4.1. The length
and diameter of the reactor were found to have an L/D ratio of 8.21, a value that confirms to the recommended
rule of thumb for plug flow reactors. The weight of the catalyst for the reactions were found to be appropriate, given
the scale of the analysis. It must be noted that a major influence on the model predictions and reaction rates is the
composition of the feed gas. In this scenario, the selected model is based on an inlet gas composition that is similar
to that produced through partially oxidised biomass-derived materials. The inlet feed was selected based on
literature findings, where partially oxidised biomass syngas was presented (Albrecht, 2004; Yusup et al., 2010; and
Alam, 2011). The input parameters and possible outlet mole fractions after equilibrium conversion as indicated by
NASA CEA programme are presented in the table below:
Table 4.2. Composition of Syngas and Outlet Mole Fractions as Presented by NASA CEA and this Model

Composition Inlet Mole Inlet Molar Flow NASA CEA: Mole Fraction PFR Model Predictions:
Fraction rate (mols-1) after Equilibrium Conversion Outlet Mole Fractions
CO 10.4 2.22×10-05 0.00333 0.0051
CO2 5.6 1.19×10-05 0.03810 0.0366
H2 83.5 1.78×10-04 0.74002 0.5789
H2O 0.5 1.07×10-06 0.04418 0.0244
CH3OH 0 0 0.17438 0.1183

38
According to the NASA CEA programme, at a temperature of 483K and pressure of 85 bars, the conversion of
CO is 96.8%, a value that was similar to that produced by this model. The programme, however indicated a
much lower equilibrium conversion of CO2, a conclusion that was also confirmed by this model. The respective
conversion of the NASA CEA and this model is presented in table 4.3:
Table 4.3. Conversion Rates of CO and CO2

Parameter CO CO2
NASA CEA: Flow Rates after Equilibrium Conversion 2.15×10-05 3.82×10-06
NASA CEA: Conversion Rates (%) 96.8 31.9
PFR Model Prediction: Conversion Rates (%) 95.3 34.7

The model-predicted conversion rates of CO and CO2 along the length of the reactor are expressed in figure 4.2.

Figure 4.2. Conversions of CO and CO2 along the Length of Reactor

It was visible that the conversions increased, as the reactions occurred across the reactor, however the as
predicted by NASA CEA (refer to table 4.3), the conversion of CO2 was much lower as compared to CO. This result
confirmed with the literature, where it has been reported that CO2 hydrogenation and conversion occurs at a much
slower rate as compared to CO (Lim et al., 2009). Yang et al (2013) reported that the formation of methanol shifts
from CO2 to CO hydrogenation reaction as the temperature in the system is lower, inferring that CO conversion
rates are greater at lower-temperature. It was also indicated in past studies, that CO2 hydrogenation is favoured
at higher temperatures due to the reverse water-gas shift reaction (Bohn, 2011). The conversions of the carbon
oxides are connected to the rate at which the reactions occur along the reactor, which includes the three main
reactions; hydrogenation of the carbon oxides and the reverse water-gas shift reaction. The reaction rates of each
main reaction in the system is presented in the figure below:

39
Figure 4.3. Rates of the Reaction along the Reactor

In the figure, rate 1 refers to the hydrogenation of CO, rate 2 as hydrogenation of CO2 and rate 3 being the reverse
water-gas shift reaction. The figure shows that the initial rates of CO and CO2 were 9.92×10-04 molkg-1s-1 and
7.64×10-04 molkg-1s-1 respectively, where a decreasing trend was observed along the length of the reactor, with
observed final rates of 3.40×10-06 molkg-1s-1 and 4.85×10-06 molkg-1s-1 for CO and CO2. The rate of the reverse
water-gas shift was found to be negative at -8.02×10-05 molkg-1s-1, indicating that the reaction is occurring in the
opposite direction. Literature has indicated that the Cu/ZnO/Al2O3 catalyst favours the water-gas shift reaction,
which results in the conversion of CO to CO2 (Yang et al., 2103). Given the occurrence of this mechanism, the
rate of the hydrogenation of CO2 in this model became greater than that of CO, as more CO2 was present in the
system. As the reactions proceeded, the rate of the reverse water-gas shift slowed, along with the hydrogenation
of the carbon oxides. According to Yang et al (2013), the presence of water and water-based intermediates in the
methanol synthesis process affects the overall kinetics and reaction mechanisms in the system, a scenario that
may have impacted the reaction rates presented in this study. In summary, it can be concluded that the
hydrogenation reaction rates decreased slowly across the length of the reactor, where they became almost
constant as thermodynamic equilibrium approached.

The rates of each reaction were compared to the weight of catalyst that was required for the desired product
composition and conversions. The results of the reaction rates and the weight of the catalyst is displayed in figure
4.4:

40
Figure 4.4. Reaction Rates versus Weight of Catalyst

The figure shows the relationship of the catalyst utilization and reaction rates of each reaction, as they occurred
along the reactor. In the computation of the total weight of catalyst required for the synthesis of methanol, the rates
of the reactions of CO and CO2 hydrogenation were major determinants in the process. The rates of the
hydrogenation reactions determine the amount of methanol produced along the reactor, a mechanism which is
affected by the occurrence of the water-gas shift reaction. Figure 4.4 shows that as the reaction reactions
proceeded along the reactor, the total weight of catalyst grew, a trend that was also observed in figure 4.3. A total
weight of 0.08kg of Cu/ZnO/Al2O3 was found to be the requirement for the reactions, given the reactor specifications
and density of the catalyst.

As the reactions occurred across the reactor, CO, CO2 and H2 were consumed producing varying amounts of
CH3OH and H2O. The mole fractions of the main reaction species along the length of the reactor is illustrated in
the figure below:

41
Figure 4.5. Mole Fraction of Reaction Species along the Length of Reactor

The exclusion of H2 in the figure was due to the range of the values; however, it was found that as CO and CO2
became hydrogenated, methanol levels rose significantly (refer to figure 4.5). It was noted that the CO was being
consumed at a greater rate than CO2, indicating that a CO-rich feed-gas maybe be favourable for methanol
synthesis. Water production was also noticeable; however, the levels were far less as compared to methanol.
According to the results displayed in figure 4.5, the production of water is mainly due to the hydrogenation of CO2.
The model indicated that the reverse water-shift gas reaction had minimal contributions to the water formation,
given that the reaction produced negative reaction rates along the reactor. The negative reaction rate implied that
the reverse water-gas shift occurred in the opposite direction, where the products, CO and H2O were being
converted to CO2 and H2. Literature has also indicated that the water–gas shift reaction is very slow, and any
contribution to water formed during this pathway is considered negligible for methanol synthesis (Graaf et al.,
1988). The results indicated that a flow stream with lowered CO2 levels may be favourable for methanol synthesis
at low temperatures given the favourable hydrogenation of CO and the reduced amounts of water in the system. It
was found that the greater the CO2 content, the higher the water production, which in practical application which
would require improved distillation and separation techniques in order to obtain high-purity methanol for the
dehydration stage. This conclusion therefore indicates that the composition of feed-gas is an important parameter
that affects product composition and possibly the operating components and costs attached with methanol
production.

4.1.3 Sensitivity Analyses of Methanol Synthesis


The model produced in the previous section was tested against various input parameters including pressure,
temperature and feed-gas composition. The simulation analyses were conducted to deduce the impacts of varied
process parameters on the output of the system and product composition. These analyses allowed for the

42
identification of the ideal conditions that can be applied for optimal production of methanol at commercial level,
with appropriate catalyst and reactor properties.
4.1.3.1 Effect of Pressure
The methanol synthesis model was first tested against pressure changes within a range of 10-85 bars in a system
at temperature of 483K. The range was selected based on the literature, where the respective conversions,
required catalyst weights and reaction lengths for each limit is shown in the table below:
Table 4.4. Effects of Pressure on Conversion Rates, Catalyst Weights and Reactor Lengths

P (bar) CO Conversion (%) CO2 Conversion (%) Catalyst (kg) Reactor Length (m)
10 36.1 -1.9 0.14 0.19
20 63.3 2.2 0.11 0.17
30 75.6 7.4 0.09 0.16
40 82.4 12.8 0.07 0.15
50 86.9 18.1 0.07 0.14
60 90.2 23.1 0.07 0.14
70 92.7 27.9 0.07 0.14
80 94.5 32.5 0.07 0.15
85 95.3 34.7 0.08 0.16

In table 4.4, it is visible that as pressure increased, the conversion rates of CO and CO2 also increased, where the
catalyst weight and reactor lengths decreased, at least up to a pressure of 85 bars. The level of conversions of CO
and CO2 are directly linked to the production of methanol in the system; figure 4.6, therefore, represents the mole
fractions of methanol along the reactors for the varying pressures of 10-80 bar:

Figure 4.6. Mole Fractions of Methanol at Varying Pressures

43
In the figure, it is clearly shown that at higher pressures, the formation of methanol in the reactor was greater, a
phenomenon that is linked to the thermodynamics of methanol synthesis. According to the Le Chatelier’s Principle,
any change in temperature, pressure or concentration to an equilibrium system would trigger a movement towards
counteracting that change (Farsi et al., 2015). Increasing the pressure in a system would result in the reactions
moving in the direction to reduce that change. For the hydrogenation reactions, one mole of CO reacts with two
moles of H2 to produce methanol, whereas for CO2, one mole of CO2 combines with three moles of H2 to produce
one mole of methanol and one mole of H2O, both reactions having greater amounts of moles as reactants. In
systems as such, an increase in pressure would shift the equilibrium to the direction where there are less moles,
in this case, to the right, where methanol formation is favoured (Sharma, 2010).
The non-equimolar and exothermic state of the hydrogenation of the carbon oxides is the main reason for the
reduction in volume of the products resulting in high-pressure favoured environments for greater product formation.

4.1.3.2 Effect of Temperature


The effects of temperature on the methanol synthesis reactions were studied in this research at a range of 483-
563K in a system at a pressure of 10 bar. The conversion rates, catalyst weights and reactor lengths for each
temperature is presented in the table below:

Table 4.5. Conversion Rates, Catalyst Weights and Reactor Lengths at Varying Temperatures

T (K) CO Conversion (%) CO2 Conversion (%) Catalyst (kg) Reactor Length (m)
483 36.1 -2.0 0.14 0.19
493 27.1 -1.5 0.07 0.15
503 18.7 -0.5 0.03 0.11
513 12.0 0.9 0.01 0.08
523 7.2 2.1 0.005 0.06
533 4.3 2.2 0.002 0.04
543 2.6 1.7 0.0009 0.03
553 1.5 1.1 0.0004 0.03
563 0.9 0.7 0.0002 0.02

In table 4.5, it is visible that the conversion rate of CO decreased as temperature increased, indicating that the
system favoured low-temperature environments. The hydrogenation reaction of CO favoured a lower temperature
for greater conversion, compared to that of CO2, where in table 4.5, it can be that the conversion of CO2 increased
with temperature up to a certain range and then decreased thereafter. Researchers have indicated that the optimal
temperature for CO2 conversion is 523-533K, given the thermodynamics of the reaction (Rahman, 2012). The
temperature analysis revealed that methanol synthesis is therefore closely connected to the syngas composition,
where the source of carbon in the methanol structure shifts from CO2 to CO as the temperature is reduced. The

44
conversion rates and methanol formation are directly linked and as such the mole fractions of methanol along the
reactor at varying temperatures were plotted to establish the effects on product formation, as shown in figure 4.7:

Figure 4.7. Mole Fractions of Methanol at Varying Temperatures

The relationship of methanol formation and temperature can be explained by the application of the Le Chatelier’s
Principle. The equations below illustrate the scenario that may occur in methanol synthesis reactors:

2H2 + CO CH3OH + Heat (R.22)


3H2 + CO2 CH3OH + H2O + Heat (R.23)
As seen in equations 22 and 23, heat becomes an added product to the hydrogenation reactions, which affects
the equilibrium state of these reactions. Following the Le Chatelier’s Principle, the system will move towards the
left to counteract the change in temperature. The methanol synthesis reactions are therefore, governed by
thermodynamics, whereby the equilibrium constant decreases in high temperature zones, reducing the ability to
produce methanol. Systems as such, therefore favours an environment with low temperatures, in order to enhance
the product formation, specifically methanol yields (Liu et al., 2003).
The results of the sensitivity analysis with respect to pressure and temperature indicated that the exothermic
hydrogenation of the carbon oxides favoured a low temperature and high pressure environment to achieve optimal
equilibrium conversions. For methanol synthesis, such conditions can result in the gases converting to liquids, a
situation that may need to be avoided for the dehydration stage of the production process. Using the Aspen plus
V8.8 software, the vapour fraction of methanol at varying temperatures and pressures were deduced. It was found
that at 483K and 85 bars, methanol has a vapour fraction of 1; however, when the pressure was increased and the
temperature lowered, the vapour fraction reduced, indicating that this temperature-pressure rate may be ideal for
the synthesis of methanol, given the required state of the products.
The conversion rates of CO and CO2 at 483K and 85 bars, along with the product state and vapour fraction of
methanol, led to the conclusion that this may be ideal operating conditions for methanol synthesis of the feed-gas
composition that was modelled.

45
4.1.3.3 Effect of Feed-Gas Composition
According to Alam (2011), the feedstock of the methanol synthesis process greatly influences the composition of
the products. In this research, the composition of the inlet gas stream was varied, in an attempt to determine the
impact of feed-gas properties on conversion rates, catalyst weights and methanol formation. The analysis was
divided into four cases, each case having a different syngas composition as presented in the table below:

Table 4.6. Cases 1-4 with the Respective Compositions

Cases 1 2 3 4
Composition CO: 10.4 CO: 5.6 CO: 24 CO: 4.6
(Mole %) CO2: 5.6 CO2: 10.4 CO2: 4 CO2: 9.4
H2: 83.5 H2: 83.5 H2: 51 H2: 65.9
H2O: 0.5 H2O: 0.5 H2O: 21 H2O: 0.04
N2: 19.56
CH3OH: 0.5

The feed-gas sensitivity analysis was conducted at operating conditions of temperature, 483K and pressure of 85
bars. For each case, the CO and CO2 conversion rates were examined and compared based on the composition
of the gas stream. The CO conversion rates for each case is displayed in figure 4.8:

Figure 4.8. CO Conversion Rates along the Reactor for Varying Feed-Gas Compositions

The figure shows that case 1 and 2, with H2/CO ratios of 8.0 and 14.9 respectively, produced greater conversions
at the selected operating conditions in the system. Case 3 having a lower H2/CO ratio of 2.1 produced lowered
rates, however, the conversions increased along the reactor. In case 3, where the water is present in the inlet
stream, the occurrence of the water-gas shift reaction may have also contributed to the CO conversion patterns.

46
With respect to case 4, with the inclusion of the inert nitrogen gas (N2), the conversion of CO was lower as
compared to the cases 1 and 2, even though the H2/CO ratio was found to be 14.3.

The conversions of CO2 were also examined with respect to each composition, where the results are presented in
the figure below:

Figure 4.9. CO2 Conversion along the Reactor for Varying Feed-Gas Compositions

The figure indicates that case 2, having greater CO2 levels as compared to case 1, produced greater conversion
rates along the reactor. Case 1 and 4, having slightly lower CO2 amounts produced lowered conversion rates,
however, the conversion of CO2 in case 3 was most interesting. Case 3, having great amounts of water in the feed
gas, resulted in negative conversion rates for CO2. The negative conversion rates in case 3 may be due to the
occurrence of the water-gas shift reaction, where CO2 was produced. The results produced by case 3 confirms
with the findings of Yang et al (2013), who indicated that the presence of water and its intermediates greatly
influence the methanol synthesis process.

The final aspect of the analysis for this section was the effect of the inlet gas composition on the weight of catalyst
and the reactor length. For each case, the required catalyst weight for the reactions along with the length of the
reactor were determined, where the results are displayed in figure 4.10:

47
Figure 4.10. Weight of Catalyst along the Reactor Length for Varying Feed-Gas Compositions

The results indicated that case 2 required the smallest reactor for the synthesis of methanol as compared to the
other cases. On the other hand, case 1 required a longer reactor, however the catalyst weight was much less than
that needed by case 2 and 3. It was interesting to note that case 4, produced results which indicated lowered
catalyst requirements for methanol formation in the reactor. The catalyst weights and reactor length are directly
connected to the rate at which the reactions occur and the composition of the inlet gas, indicating the criticality for
feed-gas analysis for optimal methanol synthesis.

In practical situations, specifically industrial applications, it is recommended to have a comprehensive analysis of


the feed-gas composition with the respective conversion rates along with the required catalyst weights and reactor
structure for such conversion. In this analysis, case 1 may have produced slightly lower conversion rates as
compared to case 2, however the catalyst requirements were much less in case 1 for such conversions. Given the
review and analysis presented above, the feed-gas composition that was selected for this study (case 1), can be
used as an appropriate inlet feed with favourable conversion rates, catalyst requirements and reactor
specifications.

4.1.4 Methanol Synthesis Model Predictions: Increased Production Level


The model presented in the section 4.1.2 was modified for increased production levels, specifically the molar flow
rates, in order to predict catalyst weights and reactor length required for increased production of methanol. This
simulation was conducted so as to determine the suitability of the kinetic model to be applied for commercial
methanol production. The inlet gas composition employed in this analysis was based on similar compositions from
partially-oxidised biomass-derived syngas, where the selected operating conditions were based on the analyses
reviewed and presented in the previous sections. The analysis of the kinetic model using the selected inputs will
allow for an understanding of its applicability for methanol synthesis in India, using the possible production pathway
as described in chapter 1.3. The simulation results are presented in table 4.7:

48
Table 4.7. Properties of the Reactor Model for Increased Methanol Production

Composition T (K) P (bar) No. of Flow Rate Catalyst Reactor Reactor


Tubes (mols-1) (kg) Diameter (m) Length (m)
CO: 10.4 483 85 5000 3100 227.48 0.28 2.09
CO2: 5.6
H2: 83.5
H2O: 0.5

The results indicated that with a molar flow rate of 3100mols-1, passing through 5000 tubes resulting in 0.62mols-1
flow within the tubes required 227.8 kg of Cu/ZnO/Al2O3 with a reactor of 2.09 m long and 0.28 m wide. The
conversion rates of the carbon oxides were similar to that presented in section 4.1.2, given similar feed-gas
composition, temperature and pressure data were used in the analysis.

4.1.5 Validation of Methanol Synthesis Model


The validation of the methanol synthesis with past studies proved difficult given the lack of relevant data in the
literature. To accurately compare the model, inlet gas compositions, temperature, pressure and outlet mole
fractions were required, much of which were not completely covered in the reported studies. However, in a study
produced by Jianqing et al (2002), the following data were presented and compared to the results of this research:

Table 4.8. Model Predictions Compared to Literature Data

Parameter T (K) P (bars) Conversion (%) Deviation (%)


Model Prediction 573 50 19.8 1.2
Jianqing et al (2002) 573 50 20.0

In the comparison analysis, the feed-gas composition was not considered, given the lack of data in the literature,
however, the model predicted conversions at the above operating conditions were based on the inlet gas stream
presented in chapter 4.1.2. The table shows that the model prediction deviates 1.2 % from that presented by
Jianqing et al (2002), inferring that the kinetic model produced in this study may be an accurate representation for
methanol synthesis.

4.2 Methanol Dehydration


4.2.1 Process Flow of Methanol Dehydration
The kinetic model for the dehydration of methanol to DME was developed for an isothermal, steady-state catalytic
fixed bed reactor. The selected system was based on the findings presented in literature and further analyses
described later in this section. In the reactor system, methanol was dehydrated to form DME and water, a reaction
that was controlled by thermodynamics. Given the limitations attached to the equilibrium conversion of methanol
to DME, a recycling unit was recommended (refer to figure 4.11). In the proposed system, the inlet gas enters a

49
mixer, which combines the feed with the recycled unreacted gases, after which it is heated to a required
temperature. The gas mixture then moves along to the reactor, which is a shell and tube system. In this reactor,
the tubes are packed with spherical Al2O3 catalyst with a void fraction of 0.4. The gas passes through the tubes at
a pressure of 2.1 bar, where the dehydration of methanol occurs on the surface of the catalyst, releasing heat in
the process. A cooling liquid is passed through the system in the shell side of the reactor, removing the excess
heat, allowing for isothermal conditions. As the reactions occur, the outlet gases pass to a distillation column,
where water and unreacted gases are separated through a two-staged distillation process. The unreacted gases
are sent to a mixer through the recycling unit. The proposed process flow with the relevant components are
displayed in the figure below:

Figure 4.11. Proposed Process Flow Sheet for Methanol Dehydration

4.2.2 Methanol Dehydration Model


The dehydration of methanol to DME was modelled using the kinetics presented by Bercic et al (1993). The
simulation of the model was conducted with the use of Microsoft Excel, where the kinetics and mathematical
equations described in chapter 3 were employed. The selected model for the dehydration process is presented in
the table below with the recommended operating conditions:

50
Table 4.9. Properties of the Recommended Methanol Dehydration Model

Property Unit Value


Inlet Temperature K 573
Pressure Bar 2.1
Number of Iterations ---- 500
Number of Tubes ---- 3000
Reactor Length m 0.48
Reactor Diameter m 0.056
Length/Diameter of Tube ---- 8.54
Catalyst Particle Diameter m 0.003
Catalyst Density kgm-3 1130
Catalyst Weight kg 1.33
Total Inlet Molar Flow Rate mols-1 0.64
Velocity of Gas Stream ms-1 1.96×10-03
Volumetric Flow Rate m3s-1 4.84×10-06

A temperature of 573K and pressure of 2.1 bar was used in the model; conditions that were selected based on
many simulation tests and analysis that were applied to the kinetics. The reactor was modelled at a lab-scale,
using a single-pass shell and tube reactor containing 3000 tubes. A total flow rate of 0.64 mols-1 was assumed,
resulting in 0.0002 mols-1 flowing through a single tube. By applying the methodologies discussed in the previous
chapter, the reactor length, catalyst weight, velocity and volumetric flow rate of the system were computed (refer
to table 4.9). The length and diameter of the reactor were found to have an L/D ratio of 8.54, a value that confirmed
to the recommended rule of thumb for plug flow reactors. The weight of the catalyst for the reactions were found
to be appropriate, given that the system was modelled at a lab-scale. It must be noted that a major influence on
the model predictions and reaction rates is the composition of the feed gas. In this scenario, the selected model is
based on an inlet gas composition where CH3OH= 1. The inlet feed was selected based on the output stream from
the methanol synthesis process, where separation techniques were employed for unreacted and unwanted gas
removal. The input parameters and possible outlet mole fractions after equilibrium conversion as indicated by
NASA CEA programme and those predicted by the model are presented in the table below:
Table 4.10. Composition of Inlet Gas and Outlet Mole Fractions Presented by NASA CEA and this Model

Composition Inlet Mole Inlet Molar Flow NASA CEA: Mole Fraction after Model Predictions: Outlet
Fraction rate (mols-1) Equilibrium Conversion Mole Fractions
CH3OH 1 0.00021 0.12774 0.12948
CH3OCH3 0 0 0.43613 0.43526
H2O 0 0 0.43613 0.21763

51
According to the NASA CEA programme, at a temperature of 573K and pressure of 2.1 bars, the conversion of
methanol is 87.2%, a value which confirmed to that produced by this model. The respective conversion of the
NASA CEA and this model is presented in table 4.11:
Table 4.11. Methanol Conversion Rates Produced by NASA CEA and this Model

Parameter CH3OH
NASA CEA: Flow Rates after Equilibrium Conversion 1.86×10-04
NASA CEA: Conversion Rates (%) 87.2
Model Prediction: Conversion Rates (%) 87.2

The conversion of methanol to DME at the selected conditions required appropriate amounts of catalyst. For
greater conversions, lower temperatures are favoured, however vast amounts of catalysts are required for such
applications. It is therefore, recommended that the system operates at 573K and 2.1 bars, where a recycling unit
is applied to improve the overall efficiency and conversions of the production process. This conclusion has also
been reported in the literature, where recycling of unreacted methanol is recommended during DME production
(Dagde et al., 2016).

The conversion of methanol to DME along the length of the reactor was examined and is displayed in figure 4.12:

Figure 4.12. Conversion of Methanol along the Length of Reactor

The figure indicates that the conversion of methanol increases across the length of the reactor, most significantly
as the reactions commenced, which then slowed across the reactor. Literature has indicated that as
thermodynamic equilibrium approaches, the reaction rates becomes slower and conversions become almost
constant (Sierra et al., 2013). The conversion of methanol is connected to the rate at which the reaction occurs
along the reactor, which includes, the overall dehydration reaction. The reaction rate for methanol dehydration
across the length of the reactor is presented in the figure below:

52
Figure 4.13. Reaction Rate of Methanol Dehydration across the Length of Reactor

The dehydration of methanol commenced at a rate of 3.38×10-04 molkg-1s-1 and concluded at 1.08×10-05 molkg-1s-1
after 500 iterations, with a final conversion of 87.2%. Comparing the conversion of methanol with the rate of the
reaction across the reactor, it is visible that as the reactions were initiated, and the rates were higher, the
conversion of methanol rose steadily; however, as the rates slowed across the reactor, the conversion of methanol
followed closely.

The rates of the reaction were compared to the weight of catalyst that was required for the desired product
composition. The results of the reaction rates and the weight of the catalyst is displayed in figure 4.14:

Figure 4.14. Reaction Rate vs Weight of the Catalyst for Methanol Dehydration

The figure displays the relationship of the catalyst utilization and reaction rates along the reactor. The reaction rate
was a major component in the computation of the weight of catalyst required for the dehydration process. Figure

53
4.14 shows that as the reaction proceeds along the length of the reactor, the total catalyst weight increased,
attaining a final value of 1.33kg for 87.2% methanol conversion. The relationship of reaction rates and catalyst
weight follows a similar trend as the reactor length, given that the two parameters are directly connected, as
discussed in chapter 3.4.

Moreover, as the reaction occurred across the reactor, CH3OH was being consumed producing varying amounts
of DME (CH3OCH3) and H2O. The mole fractions of the main reaction species along the length of the reactor is
illustrated in the figure below:

Figure 4.15. Mole Fractions of the Main Reaction Species along the Reactor

The figure shows that as methanol is dehydrated, DME and water is produced along the reactor. Figure 4.15
indicated that methanol concentration decreased as the reactor length increased, while the levels of DME and
water increased. It is visible that that as the reaction commenced, methanol was dehydrated at a rapid rate, while
DME and water production increased steadily; however, as the dehydration slowed across the reactor, the
production levels became almost constant. At the selected operating conditions, 87.2% of methanol was converted
to water and DME. Additionally, it can be seen that the production of DME is much greater than that of water, a
phenomenon which is connected to the application of the Al2O3 catalyst. Literature has indicated that the Al2O3
catalyst favours the production of DME, a major consideration for catalyst selection (Dagde et al., 2016). The
reduction of water in the product composition is favoured, given reduced costs for distillation and high-purity level
DME production.

4.2.3 Sensitivity Analyses of Methanol Dehydration


The model produced for the dehydration of methanol to DME was tested against various input parameters including
pressure, temperature and feed-gas composition. The simulation analyses were conducted to determine the

54
impacts of varied process parameters on the output of the system and the product composition. These analyses
allowed for the identification of the ideal conditions that can be applied for optimal DME production at industrial
levels, with appropriate catalyst and reactor properties.
4.2.3.1 Effect of Pressure
The review and examination of the thermodynamics and rate equation of the methanol dehydration process
revealed that pressure has no effect on equilibrium conversion of methanol and reaction rates. In the rate equation
(refer to equation 3.14), it can be seen that pressure variations were excluded in the kinetics presented by Bercic
et al (1993), thereby allowing the model to be independent of pressure effects. Likewise, analysis of the chemical
equation and thermodynamics of the methanol dehydration reaction indicated that varying pressure in the reactor
has no effect on the conversion rates and production of DME. In the dehydration of methanol to DME, two moles
of methanol produce one mole of DME and one mole of water, making the number of moles of reactants and
products equal. The equimolar state of the reaction, therefore allows the equilibrium of the system to be
independent of pressure. According to Sharma (2010), in an equimolar reaction, any change in pressure or volume
does affect the state of the equilibrium, allowing the reaction to proceed in either direction.

The equimolar state of the methanol dehydration reaction, therefore allows the reactor system to operate at lower
pressures with no impact on product composition. For the reasons outlined above, the effect of pressure on the
system were excluded in this report, however simulation works were conducted and confirmed to that presented
in literature.

4.2.3.2 Effect of Temperature


The effects of temperature on the methanol dehydration reaction was examined in this study at a range of 473-
613K in a system at a pressure of 2.1 bars. The conversion rates, catalyst weights and reactor lengths for each
temperature is presented in the table below:

Table 4.12. Conversion Rates, Catalyst Weights and Reactor Lengths at Varying Temperatures

Temperature (K) Conversion (%) Catalyst (kg) Reactor Length (m)


473 91.7 62.64 1.76
493 90.8 25.50 1.24
513 89.9 11.17 0.95
533 89.0 5.20 0.76
553 88.1 2.57 0.61
573 87.2 1.33 0.48
593 86.4 0.72 0.35
613 85.5 0.41 0.30

Table 4.12 indicates that the conversion rate of methanol to DME decreased as temperature increased, indicating
that the system favoured low-temperature environments. However, it is noticeable that at lower temperatures, vast

55
amounts of catalyst are required to enhance the reaction, a situation which may be impractical for industrial
production of DME. Moreover, the conversion rates and DME formation are directly connected and as such the
mole fractions of DME along the reactor at varying temperatures were plotted to establish the effects on product
formation, as shown in figure 4.16:

Figure 4.16. Mole Fractions of DME at Varying Temperatures

The relationship of methanol dehydration and temperature was explored using the Le Chatelier’s Principle. The
equation below illustrates the scenario that may occur during methanol dehydration:

2CH3OH CH3OCH3 + H2O + Heat (R.24)


As expressed in the equation above, heat becomes an added product to the reaction, which affects the equilibrium
state of the reaction. Guided by the Le Chatelier’s Principle, the system will move towards the left to counteract
the increase in temperature. Methanol dehydration reaction is therefore limited by thermodynamics, whereby the
equilibrium constant decreases in high temperature zones, reducing the ability to form DME. Reactor systems with
such characteristics favours an environment with low temperatures, in order to enhance the product formation,
specifically DME yields (Dagde et al., 2016). According to Farsi et al (2011), high conversion of methanol to DME
have been reported for the temperature range of 473-623K, after which the conversion rates becomes almost
constant above this range. Literature has reported that around 570-590K, the methanol dehydration system
approaches equilibrium. Additionally, catalyst deactivation at higher temperatures were also considered in the
temperature analysis, given that the Al2O3 catalyst is crucial for DME selectivity.
Given the analysis presented in this section, a temperature of 573K was selected based on the conversion rate,
catalyst weight and reactor length.
4.2.3.3 Effect of Feed-Gas Composition
Literature has shown that the composition of the inlet gas greatly affects the conversions rates of methanol and
the production levels of DME in the dehydration process (Farsi, 2015). In this study, the composition of the inlet

56
gas stream was varied, in an attempt to determine the impact of feed-gas properties on conversion rates, catalyst
weights, reactor lengths and DME formation. The analysis was divided into three cases, each case having a
different feed-gas composition as presented in the table below:

Table 4.13. Cases 1-3 with the Respective Compositions

Case Composition % Weight of Catalyst (kg) Reactor Length (m)


1 CH3OH: 100 1.33 0.48
2 CH3OH: 90 1.32 0.46
H2O: 10
3 CH3OH: 85 1.29 0.45
H2O:10
CH4: 5

The feed-gas sensitivity analysis was conducted at operating conditions of temperature, 573K and pressure of 2.1
bars. For each case, the methanol conversion rates were examined and compared based on the composition of
the gas stream. The conversion rates along the length of the reactor is represented in the figure below:

Figure 4.17. Conversion of Methanol for Varying Feed-Gas Compositions

The figure indicates that the inlet gas which contained methanol in a mole fraction of one, resulted in the greatest
conversion along the reactor. It is quite visible from the figure, that as the mole fraction of methanol decreased in
the feed-gas, so did the conversion rates across the reactor, where case 2 resulted in higher conversion as
compared to case 3. As shown in table 4.13, the greater the methanol composition in the inlet gas stream, the
greater the catalyst requirements and reactor length for the dehydration reaction, given that the conversion rates
would be also be higher. Moreover, higher catalyst requirements imply that a longer reactor would be required for

57
the dehydration reaction. Literature has confirmed the results reported in this analysis, where Farsi (2015) indicated
that the presence of water vapour in the reactor system decreases the equilibrium conversion of methanol and
noted that water removal in the first production stage, that is, methanol synthesis would result in favourable DME
production. The analysis, however in this study indicated that both the unreacted methane and water in the feed-
gas reduced the equilibrium conversion of methanol in the reactor and therefore, an inlet gas with composition of
CH3OH=1 is recommended for the second stage of the DME production.

4.2.4 Dehydration Model Predictions: Increased Production Level


The model presented in the section 4.2.2 was adjusted to account for increased production levels of DME, where
the molar flow rates were adjusted in order to predict the required catalyst weight and reactor length. This
simulation was conducted so as to determine the suitability of the kinetic model to be applied for commercial DME
production. The inlet gas composition employed in this analysis was based on the outputs described in the
methanol synthesis presented earlier, where the selected operating conditions were based on the analyses
presented in the previous sections. The analysis of the kinetic model using the selected inputs will allow for an
understanding of its applicability for DME production in India, using the possible production pathway as described
in chapter 1.3. The simulation results are presented in table 4.14:

Table 4.14. Properties of the Reactor Model for Increased DME Production

Composition Temp. Pressure No. of Flow Rate Weight of Reactor Reactor


% (K) (bar) Tubes (mols-1) Catalyst (kg) Diameter (m) Length (m)
CH3OH: 100 573 2.1 5000 750 936.38 0.53 3.76

The total molar flow rate for the commercial production was assumed to be 750 mols-1, flowing through 5000 tubes
in a catalytic shell and tube reactor. Therefore, 0.15 mols-1 was determined as the molar rate flowing through a
single tube, requiring a catalyst weight of 936.38 kg along a reactor of 3.76 m with a diameter of 0.53 m for 87.2%
methanol conversion.

The simulation indicates that the kinetic model for methanol dehydration can be applied for varying flow rates of
inlet gas stream, given the required DME production levels.

4.2.5 Validation of Methanol Dehydration Model


The dehydration model produced in this study was compared to the results presented in past research. Due to the
lack of sufficient data in the literature, the conversion rates of methanol at specific operating conditions were the
only parameters that were examined. In the table below, the conversion rates of methanol as reported by Dagde
et al (2016), is compared against the results produced in this study:

58
Table 4.15. Comparison of Model Predictions with Literature Data

Parameters Composition T (K) P (bars) Conv. rates (%) Deviation (%)


PFR Model Prediction CH3OH=1 533 18.2 88.9
1.3
Dadge et al (2016) CH3OH=1 533 18.2 87.9

The data from the literature was sourced from a study produced by Dagde et al in 2016, where DME production
using Al2O3 catalyst was studied. The predictions from this study shows a 1.3% deviation from that presented in
the literature, a value that infers minimal difference between the models produced in the two studies. The
comparison therefore, indicated that this model is reasonably accurate with respect to its application in methanol
dehydration systems, given the validation with past studies.

4.3 Conclusion
The review and examination of the methanol synthesis model revealed that a CO-rich feed gas with operating
conditions of 483K and 85 bar resulted in optimal conversions of CO and CO2. The dehydration of methanol to
DME was also modelled using similar methodologies, which indicated greater conversions were achieved at higher
methanol concentrations with operating conditions of 573K. The methanol dehydration model was, however, found
to be independent of pressure, given the reaction rates applied and the equimolar state of the chemical reaction.

59
Chapter 5: Conclusion & Recommendations for Further Work
5.1 Conclusion
This study investigated the conversion of biomass-derived syngas to produce an alternative diesel fuel for India.
DME was found to be an ideal substitution for diesel fuel, given the necessity for cleaner, renewable and
sustainable liquid transport fuels in India. The increasing population coupled with an emerging economy has
triggered many concerns relating to global warming and energy security. The country has since realized the
urgency for renewable energy forms, with the utilization of indigenous and inexpensive resources, specifically
agricultural waste. The aim of this study was to design an ideal gas reactor for the conversion of biomass-derived
syngas to DME, with the application of the reaction pathways, kinetic models and mathematical equations
presented in past studies.
The production of DME was modelled using the two-staged process where methanol synthesis and dehydration
were examined. The study indicated that the methanol synthesis process employs three overall reactions, including
the hydrogenation of CO and CO2 and the reverse water-gas shift reaction. It was found that the hydrogenation of
CO occurred at the greatest rate as compared to the other reactions, and favoured much lower temperatures than
CO2. The rate of the reverse water-gas shift reaction was reported to be negative, indicating that the reaction was
moving in the opposite direction. The review of the synthesis model also indicated the thermodynamic limitations
of the chemical reactions, whereby the exothermicity and non-equimolar state led to favourable conversion rates
at lower temperatures and higher pressures. Moreover, it was indicated that the feed-gas composition also greatly
affects the conversion rates of CO and CO2, indicating the need for a comprehensive analysis of the input
parameters for the production process. The selection of the ideal model and operating conditions were therefore
based on the simulation studies and revealed that optimal equilibrium conversion and production of methanol
occurred at 483K and 85 bar, given appropriate catalyst requirements, reactor specifications and product
composition.
The second stage of the conversion process was investigated with much emphasis of the dehydration reaction of
methanol to produce DME. The analysis indicated that the production of DME is the greatest with increased
methanol composition in the feed-gas, however the catalyst and reactor requirements were slightly higher as
compared to other inlet gas ratios. The equimolar chemical reaction resulted in the catalytic dehydration process
to be independent of pressure changes, allowing the system to be operated at low pressures. The examination of
the reaction kinetics also indicated that the reaction rates were unaffected by pressure, given the exclusion of
partial pressures in the equation. Additionally, it was found that the dehydration process was also restricted by
thermodynamics, whereby the exothermic reaction indicated greater methanol conversions at lower temperatures.
However, the simulation analysis revealed that for very low temperatures, massive amounts of catalyst will be
required, resulting in the selection of 573K as the ideal temperature for the reaction, given the catalyst and reactor
specifications.
The models presented in this study for methanol synthesis and dehydration produced reasonably accurate results,
given the comparisons made with the literature. Each model was validated against past studies, where

60
approximately 1.0% deviation were reported. The models were also altered for increased productions similar to
that at commercial level, where favourable results were reported. The kinetic models for the synthesis and
dehydration of methanol to DME investigated in this study, can be therefore be applied to simulate the conversion
of biomass-derived syngas in a catalytic, shell and tube fixed-bed reactor to produce DME, as an alternative diesel
fuel for India.
5.2 Future Work
In this study, the synthesis and dehydration of methanol to produce DME were successfully modelled using the
kinetics provided in the literature. The models were analysed and reviewed for various parameters and effects,
resulting in a reasonably accurate representation for the production process. This study can therefore, be used to
conduct further simulation analysis with the inclusion of other parameters such as pressure drop within the reactor,
which would allow to have a more comprehensive understanding of the reaction rates and the impact on product
composition. The model produced in this study was based on isothermal conditions, given the high sensitivity of
the conversion rates to temperature. In the future, reactor modelling for methanol synthesis and dehydration can
be done using membrane reactors, which have been reported to have enhanced temperature control.

61
References
Akpan, U.F., and Akpan, G.E. 2012. The contribution of energy consumption to climate change: A feasibility
policy direction. International Journal of Energy Economics and Policy. [Online]. 2(1), pp.21-33. [Accessed 19
March 2017]. Available from: https://www.econjournals.com/index.php/ijeep/article/download/96/67

Alam, M. 2011. Kinetics and deactivation in the methanol synthesis reaction. Master’s Thesis. Norwegian
University of Science and Technology. [Online]. [Accessed 26 May 2017]. Available from:
https://brage.bibsys.no/xmlui/bitstream/id/388283/6654_FULLTEXT.pdf

Albrecht, B.A., Reactor modelling and process analysis for partial oxidation of natural gas, PhD Thesis,
University of Twente, 2004. [Online]. [Accessed 08 March 2017]. Available from:
https://www.utwente.nl/en/et/thw1/Publications/PhD-theses/Albrecht.pdf

Arthur, T. 2010. Control structure design for methanol process. Master’s Thesis. Norwegian University of Science
and Technology. [Online]. [Accessed 26 May 2017]. Available from:
https://www.researchgate.net/file.PostFileLoader.html?id...assetKey...

Askgaard, T.S., Norskov, J.K., Ovensen, C.V., and Stoltze, P. 1995. A kinetic model of methanol synthesis.
Journal of Catalysis. [Online]. 156(1995), pp.229-249. [Accessed 03 July 2017]. Available from:
http://www.sciencedirect.com/science/article/pii/S002195178571250X

Azizi, Z., Rezaemanesh, M., Tohidian, T., Rahimpour, M.R. 2014.Dimethyl ether: a review of technology and
production challenges. Chemical Engineering and Processing. [Online]. 82(2014), pp.150-172. [Accessed 12
July 2017]. Available from:
https://www.researchgate.net/publication/263283870_Dimethyl_Ether_A_Review_of_Technologies_and_Product
ion_Challenges

Bakhtiary, H., and Hayer, F. 2008. Kinetics and reactor modelling of methanol synthesis from synthesis gas.
COMSOL Conference, November 06, Hanover. [Online]. [Accessed 24 April 2017]. Available from:
https://www.comsol.com/paper/kinetics-and-reactor-modeling-of-methanol-synthesis-from-synthesis-gas-5655

Bandyopadhyal, K.R.and Roy, S. 2015. Biofuel promotion in India for transport: exploring the grey areas.
[Online]. New Delhi: Energy and Resources Insitute. [Accessed 02 March 2017]. Available from:
http://www.teriin.org/policybrief/docs/biofuel.pdf

Bansal, G.and Bandivadekar, A.2013. Overview of India’s vehicle emissions control program, past successes
and future prospects. [Online]. Washington: International Council on Clean Transportation. [Accessed 12 March
2017]. Available from: http://www.theicct.org/sites/default/files/publications/ICCT_IndiaRetrospective_2013.pdf

Basavaraj, G., Rao, P.P., Reddy, C.R., Kumar, A.A., Rao, P.S., and Reddy, B.V.S. 2012. A review of the national
biofuel policy in India: a critique of the need to promote alternative feedstocks. [Online]. India: International Crop
Research institute for the Semi-Arid Tropics. [Accessed 05 July 2017]. Available from:
http://exploreit.icrisat.org/sites/default/files/uploads/1379310575_1378353812_WPS_34(1)[1].pdf

Baskaran, R. 2015. Analysis on synthesis, storage and combustion of DME as fuel in CI engines. International
Journal for Research in Applied Science and Engineering Technology. [Online]. 3(1), pp.133-140. [Accessed 17
April 2017]. Available from: http://www.ijraset.com/fileserve.php?FID=1542

62
Bohn, K. 2011. Design configurations of the methanol synthesis loop. Master’s Thesis. Norwegian University of
Science and Technology. [Online]. [Accessed 28 May 2017]. Available from:
https://brage.bibsys.no/xmlui/bitstream/handle/11250/2366801/6421_FULLTEXT.pdf?sequence=1

Bourg, H.M. 2006. Future perspective of DME. [Online]. France. [Accessed 12 March 2017]. Available from:
http://www.aboutdme.org/EFIClient/files/ccLibraryFiles/Filename/000000000735/2006_WorldGasConf_deMestier
duBourg_Total.pdf

Bercic, G., and Levec, J. 1993. Catalytic dehydration of methanol to dimethyl ether. A kinetic investigation and
reactor simulation. Industrial Engineering Chemistry Research. [Online]. 1993(32), pp.2478-2484. [Accessed 02
June 2017]. Available from: http://pubs.acs.org/doi/abs/10.1021/ie00023a006

Bussche, K.M.V., and Froment, G.F. 1996. A steady-state kinetic model for methanol synthesis and the water-
gas shift reaction on a commercial Cu/ZnO/Al2O3 catalyst. Journal of Catalysis. [Online]. 161(1996), pp.1-10.
[Accessed 10 July 2017]. Available from: https://terpconnect.umd.edu/~nsw/chbe446/VandenBussche-
syngas2MeOH.pdf

Dadge, K.K. and Slyvanus, H.U.2016. Modelling of a tubular fixed-bed reactor for the production of dimethyl
ether using alumina catalyst. International Journal of Chemical and Process Engineering Research.[Online]. 3(2),
pp.23-34. [Accessed 18 April 2017]. Available from: http://www.pakinsight.com/pdf-files/eng/65/IJCPER-2016-
3(2)-23-34.pdf

de la Rue du Can, S., Letschert, V, McNeil, M., Zhou, N. and Sathaye, J. 2009. Residential and transport energy
use in India: past trend and future outlook. [Online]. United States: Ernest Orlando Lawrence Berkeley National
Laboratory. [Accessed 10 March 2017]. Available from: http://re.indiaenvironmentportal.org.in/files/LBNL-
1753E.pdf

De Maria, R., Diaz, I., Rodriguez, M.and Saiz, A. 2013. Industrial methanol from syngas: kinetic study and
process simulation. International Journal of Chemical Reactor Engineering. [Online]. 11(1), pp.469-477.
[Accessed 25 April 2017]. Available from: https://www.degruyter.com/downloadpdf/j/ijcre.2013.11.issue-1/ijcre-
2013-0061/ijcre-2013-0061.pdf

European Business and Technology Centre.2013. Transport and energy in India. [Online]. New Delhi: European
Business and Technology Centre. [Accessed 04 April 2017]. Available from:
http://ebtc.eu/pdf/130925_REP_Report-on-Transport-and-Energy-in-India_Web.pdf

European Biofuels Technology Platform. 2016. Biofuel fact sheet. [Online]. European Biofuels Technology
Platform. [Accessed 27 may 2017]. Available from:
http://www.etipbioenergy.eu/?option=com_content&view=article&id=180

Farsi, M., Eslamloueyan, R., and jahanmiri, A. 2011. Modelling, simulation and control of dimethyl ether
synthesis in an industrial fixed-bed reactor. Chemical Engineering and Processing. [Online]. 50(2011), pp.85-94.
[Accessed 05 June 2017]. Available from:
https://www.researchgate.net/publication/251538888_Modeling_simulation_and_control_of_dimethyl_ether_synt
hesis_in_an_industrial_fixed-bed_reactor

63
Farsi, M. 2015. Dynamic modelling and controllability analysis of DME production in an isothermal fixed bed
reactor. Chemical Engineering Research Bulletin. [Online]. 17(2015), pp.40-51. [Accessed 05 June 2017].
Available from: www.banglajol.info/index.php/CERB/article/download/22917/15770

Ghosh, S.K. 2016. Potential of economic utilization of biomass waste in India- Implications towards the SDG’s.
In: Seventh Regional 3R Forum in Asia and the Pacific, 2-4 November 2016, Australia. [Online]. Australia: United
Nations Centre for Regional Development, pp.1-22. [Accessed 16 April 2017]. Available from:
http://www.uncrd.or.jp/content/documents/4438Combined%20FrontPage&Background-Sadhan%20Ghosh-PS-
3.pdf

Graaf, G.H., Stamhuis, E.J., Beenackers, A.C.C.M. 1988. Kinetics of low-pressure methanol synthesis. Chemical
Engineering Science. [Online]. 43(12), pp.3185-3195. [Accessed 15 May 2017]. Available from:
http://www.sciencedirect.com/science/article/pii/0009250988851273

Hansen, J., Kharecha, P., Sato, M., Masson-Delmotte, V., Ackerman, F., Beerling, D.J., Hearty, P.J., Hoegh-
Guldberg, O., Hsu, S.L., Parmesan, C., Rockstrom, J., Rohling, E.J., Sachs, J., Smith, P., Steffen, K., Susteren,
L.V., Schuckman, K.V., and Zachos, J.C. 2013. Assessing “dangerous climate change”: Required reduction of
carbon emissions to protect young people, future generations and nature. PLOS one. [Online]. 8(12), pp.1-26.
[Accessed 27 April 2017]. Available from: https://www.ncbi.nlm.nih.gov/pubmed/24312568

Hu, J., Yu, F.and Lu, Y.2012. Application of Fischer-Tropsch synthesis in biomass to liquid conversion. Catalysts.
[Online]. 2012(2), pp.303-326. [Accessed 01 March 2017]. Available from:
https://pdfs.semanticscholar.org/be44/8b0f812cfa2eee81374d8f76310a2d97f1d0.pdf

International Energy Agency.2011. Energy transition for industry: India and the global context. [Online]. Paris:
International Energy Agency. [Accessed 03 April 2017]. Available from:
https://www.iea.org/publications/freepublications/publication/india_industry_transition_28feb11.pdf

International Energy Agency. 2015. India Energy Outlook. [Online]. Paris: International Energy Agency.
[Accessed 15 March 2017]. Available from:
https://www.iea.org/publications/freepublications/publication/IndiaEnergyOutlook_WEO2015.pdf

IPCC, 2014: Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth
Assessment Report of the Intergovernmental Panel on Climate Change [Core Writing Team, R.K. Pachauri and
L.A. Meyer (eds.)]. IPCC, Geneva, Switzerland. [Online]. [Accessed 24 June 2017]. Available from:
https://www.ipcc.ch/report/ar5/syr/

Jalil, R., Miyazawa, T.,Sakanishi, K., Yusuff, M.N.M., Ibrahim, W.A. 2010. Synthesis of dimethyl ether (DME)
from woody biomass syngas via one step synthesis process over bifunctional catalysts. In: 7th Biomas Asia
Workshop, 29 November- 01 Dcember 2010, Indonesia. [Online]. [Accessed 06 march 2017]. Available from:
https://www.researchgate.net/publication/267864617_Synthesis_of_Dimethyl_Ether_DME_from_Woody_Biomas
s_Syngas_via_One_Step_Synthesis_Process_over_Bi-functional_Catalysts?enrichId=rgreq-
048f77d110b286fc73adcfb75fa8dcc8-
XXX&enrichSource=Y292ZXJQYWdlOzI2Nzg2NDYxNztBUzoyODU4NTQ5MzAwOTYxMjhAMTQ0NTE2NDUzN
DAxNg%3D%3D&el=1_x_2&_esc=publicationCoverPdf

64
Jianqing, Z., Noritatsu, T., Kaoru, F. 2002. New process of low-temperature methanol synthesis from CO/CO2/H2
based on dual-catalysis. Science in China. [Online]. 45(1), pp. 106-112. [Accessed 16 August 2017]. Available
from: http://chem.scichina.com:8081/sciBe/fileup/PDF/02yb0106.pdf

Kittleson, D., Watts, W., Bennett, D., Taff, S., and Chan, C. 2010. Performance and emissions of a second-
generation biofuel- DME. Initiative for Renewable Energy and the Environment, 30 November, St. Paul
Minnesota. [Online]. [Accessed 12 May 2017]. Available from:
http://www.me.umn.edu/centers/cdr/reports/E3_Kittelson.pdf

Kubota, T., Hayakawa, I., Mabuse, H., Mori, K., Ushikoshi, K., Watanabe, T., and Saito, M. 2001. Kinetic study of
methanol synthesis from carbon dioxide and hydrogen. Applied Organometallic Chemistry. [Online]. 2001(15),
pp.121-126. [Accessed 26 July 2017]. Available from: http://onlinelibrary.wiley.com/doi/10.1002/1099-
0739(200102)15:2%3C121::AID-AOC106%3E3.0.CO;2-3/pdf

Laurence, T., Thermo-economic evaluation of the production of liquid fuels from biomass, MSc Thesis,Swiss
Federal Institute of Technology, 2009. [Online]. [Accessed 12 March 2017]. Available from:
https://infoscience.epfl.ch/record/142291/files/PDM_LaurenceTock_02_2009.pdf

Lim, H.W., Park, M.J., Kang, S.H., Chae, H.J., Bae, J.W., and Jun, K.W. 2009. Modelling of the kinetics for
methanol synthesis using Cu/ZnO/Al2O3/ZrO2 catalyst: influence of carbon dioxide during hydrogenation.
Industrial and Engineering Chemistry Research. [Online]. 2009(48), pp.10448-10455. [Accessed 22 July 2017].
Available from: http://pubs.acs.org/doi/abs/10.1021/ie901081f

Liu, X.M., Lu, G.Q., Yan, Z.F., and Beltramini, J. 2003. Recent advances in catalysts for methanol synthesis via
hydrogenation of CO and CO2. Industrial Chemical Engineering Research. [Online]. 2003(42), pp.6518-6530.
[Accessed 05 June 2017]. Available from: http://pubs.acs.org/doi/abs/10.1021/ie020979s

Maji, S., Ahmed, S., Siddiqui, W.A., Aggarwal, S. and Kumar, A.2015. Impact of di-methyl ether (DME) as an
additive fuel for compression ignition engine in reduction of urban air pollution. American Journal of
Environmental Protection. [Online]. 3(2015), pp.48-52. [Accessed 01 March 2017]. Available from:
http://pubs.sciepub.com/env/3/2/2/

Mayra, O., and Leiviska, K. 2008. Modelling in methanol synthesis. [Online]. Finland: University of Oulu.
[Accessed 03 June 2017]. Available from: http://jultika.oulu.fi/files/isbn9789514290145.pdf

Ministry of Statistics and Programme Implementation.2015. Statistics related to climate change-India 2015.
[Online]. New Delhi: Ministry of Statistics and Programme Implementation. [Accessed 04 April 2017]. Available
from: http://www.mospi.gov.in/sites/default/files/publication_reports/climateChangeStat2015.pdf

Nahar, G., Mote, D. and Dupont, V.2017. Hydrogen production from reforming of biogas: review of technological
advances and an Indian perspective. Renewable and Sustainable Energy Reviews. [Online]. 76(2017), pp.1032-
1052. [Accessed 22 March 2017]. Available from: http://ac.els-cdn.com/S1364032117302502/1-s2.0-
S1364032117302502-main.pdf?_tid=a0fdc300-2c15-11e7-8911-
00000aacb35d&acdnat=1493385886_b709c4d32ca0c37aed0add871cfb401c

National Aeronautics and Space Administration. 2016. Global climate change. [Online]. [Accessed 03 April
2017]. Available from: https://climate.nasa.gov/vital-signs/global-temperature/

65
Ogawa, T., Inoue, N., Shikada, T., and Ohno, Y. 2003. Direct dimethyl ether synthesis. Journal of Natural Gas
Chemistry. [Online]. 12(2003), pp.219-227. [Accessed 25 April 2017]. Available from:
http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.585.7743&rep=rep1&type=pdf

Ohno, Y., Yagi, H., Inoue, N., Okuyama, K.and Aoki, S. 2005. Slurry phase DME direct synthesis technology-
100 tons/day demonstration plant operation and scale-up study. International Journal of Chemical Reactor
Engineering. [Online]. 8(1), pp.1-12. [Accessed 22 March 2017]. Available from:
http://www.sciencedirect.com/science/article/pii/S0167299107801650

Park, Y.S., Kim, H.S, Shun, D., Song, K.S., and Kang, S.K. 2000. Attrition characteristics of alumina catalyst for
fluidised bed incinerator. Korean Journal of Chemical Engineering. [Online]. 17(3), pp.284-287. [Accessed 21
August 2017]. Available from: https://link.springer.com/article/10.1007/BF02699041

Puladian, N., Li, J., and Pang, S. 2013. Development of an integrated system model for the production of
Fischer-Tropsch liquid fuels from biomass. [Online]. [Accessed 09 May 2017]. Available from:
http://www.conference.net.au/chemeca2013/papers/27002.pdf

Rahman, D. 2012. Kinetic modelling of methanol synthesis from carbon monoxide, carbon dioxide and hydrogen
over Cu/ZnO/Cr2O3 catalyst. Master’s Thesis. San Jose State University. [Online]. [Accessed 29 May 2017].
Available from: http://scholarworks.sjsu.edu/cgi/viewcontent.cgi?article=7695&context=etd_theses

Roh, H.G., Lee, D.and Lee, C.S.2015. Impact of DME-biodiesel, diesel-biodiesel and diesel fuels on the
combustion and emission reduction characteristics of a CI engine according to pilot and single injection
strategies. Journal of the Energy Institute. [Online]. 88(2015), pp.3376-385. [Accessed 05 March 2017]. Available
from: https://www.researchgate.net/publication/269729788_Impact_of_DME-biodiesel_diesel-
biodiesel_and_diesel_fuels_on_the_combustion_and_emission_reduction_characteristics_of_a_CI_engine_acco
rding_to_pilot_and_single_injection_strategies

Scott, A.and Seth, P. 2013. The political economy of electricity distribution in developing countries. [Online].
London: UKAid. [Accessed 13 March 2017]. Available from: https://www.odi.org/sites/odi.org.uk/files/odi-
assets/publications-opinion-files/8332.pdf

Sharma, B.K. 2010. Krishna’s objective question bank in chemistry. [Online]. India: Krishna Prakashan Media.
[Accessed 27 July 2017]. Available from: https://books.google.co.uk/books?id=YyoUNYgBYJoC&pg=SL16-
PA366&lpg=SL16-
PA366&dq=why+do+exothermic+reactions+favour+lower+temperatures?&source=bl&ots=OVz7zy2vzk&sig=_Ic
BNpl03l92RZ8-
A9_DQwgZgnc&hl=en&sa=X&ved=0ahUKEwiBrb7p58rVAhUEIcAKHb1jCN04ChDoAQgwMAI#v=onepage&q=w
hy%20do%20exothermic%20reactions%20favour%20lower%20temperatures%3F&f=false

Siedlecki, M., de Jong, W., Verkooijen, A.H.M. 2011. Fluidised bed gasification as a mature and reliable
technology for the production of bio-syngas and applied in the transportation of liquid transportation fuels- a
review. Energies. [Online]. 2011(4), pp.389-434. [Accessed 24 April 2017]. Available from:
http://www.mdpi.com/1996-1073/4/3/389

Sierra, I., Erena, J., Aguayo, A.T., Ateka, A., and Bilbao, J. 2013. Kinetic modelling of dehydration of methanol to
dimethyl ether over y-Al2O3 catalyst. Chemical Engineering Transactions. [Online]. 32, pp.613-618. [Accessed
18 June 2017]. Available from: http://www.aidic.it/cet/13/32/103.pdf

66
Szybist, J.P., McLaughlin, S.and Iyer, S. 2014. Emissions and performance benchmarking of a prototype
dimethyl ether-fuelled heavy-duty truck. [Online]. Tennesse: Oak Ridge National Laboratory. [Accessed 10
March 2017]. Available from: http://www.afdc.energy.gov/uploads/publication/ornl_dme_tm-2014-59.pdf

Taupy. J.A. no date. DME, a new energy carrier. [Online]. [Accessed 24 March 2017]. Available from:
http://siteresources.worldbank.org/EXTGGFR/Resources/578068-1250182538576/taupy.pdf

United Nations Development Programme. 2016. Sustainable development goals. [Online]. New York: United
Nations Development Programme. [Accessed 02 March 2017]. Available from:
http://www.un.org/sustainabledevelopment/energy/

World Energy Council. 2011. Global transport scenarios 2050. [Online]. London: World Energy Council.
[Accessed 05 March 2017]. Available from: https://www.worldenergy.org/wp-
content/uploads/2012/09/wec_transport_scenarios_2050.pdf

Villa, P., Forzatti, P., Buzzi-Ferraras, G., Garone, G., and Pasquon, I. 1985.Synthesis of alcohols from carbon
oxides and hydrogen 1. Kinetics of the low-pressure methanol synthesis. Industrial Engineering Chemistry
Process Development. [Online]. 24(1), pp.12-19. [Accessed 09 June 2017]. Available from:
http://pubs.acs.org/doi/abs/10.1021/i200028a003

Wilkison, S.K. Reaction kinetics in formulated industrial catalysts.2014. PhD Thesis, University of Birmingham.
[Online]. [Accessed 17 June 2017]. Available from: http://etheses.bham.ac.uk/5113/1/Wilkinson14EngD.pdf

World Bank. 2017. Energy use (kg of oil per capita). [Online]. [Accessed 03 April 2017]. Available from:
http://data.worldbank.org/indicator/EG.USE.PCAP.KG.OE?locations=IN

Wyoming, L.2014. Development of a novel helical channel reactor for syngas conversion. [Online]. Wyoming:
Ambre Fuels. [Accessed 06 March 2017]. Available from: https://www.uwyo.edu/ser/_files/docs/cleancoal/2014-
presentations/drinnan_nicholas.pdf

Xiao, K., Wang, Q., Qi, X., and Zhong, L. 2017. For better industrial Cu/ZnO/Al2O3 methanol synthesis catalyst: A
compositional study. Catalysis Letters. [Online]. 2017(147), pp.1581-1591. [Accessed 08 August 2017]. Available
from: https://link.springer.com/article/10.1007/s10562-017-2022-8

Xie, Q., Chen, P., Peng, P., Lui, S., Peng, P., Zhang, B., Cheng, Y., Wan, Y., Liu, Y.and Ruan, R. 2015. Single-
step synthesis of DME from syngas on CUZnAl-zeolite bifunctional catalysts: the influence of zeolite type. Royal
Society of Chemistry. [Online]. 2015(5), pp.26301-26307. [Accessed 02 March 2017]. Available from:
http://pubs.rsc.org/en/Content/ArticleLanding/2015/RA/c5ra02814k#!divAbstract

Yang, Y., Mims, C.A., Mei, D.H., Peden, C.H.F., and Campbell, C.T. 2013. Mechanistic studies of methanol
synthesis over Cu from CO/CO2/H2/H2O mixtures: the source of C in methanol and role of water. Journal of
Catalysis. [Online]. 298(2013), pp.10-17. [Accessed 28 May 2017]. Available from:
http://depts.washington.edu/campbelc/pdf/MeOH%20synthesis%20C%20source%20H2O%20role.pdf
Yusup, S., Anh, N.P., and Zabiri, H. 2010. A simulation study of an industrial methanol reactor model based on
simplified steady-state model. IJRRAS. [Online]. 5(3), pp.213-222. [Accessed 03 June 2017]. Available from:
http://www.arpapress.com/volumes/vol5issue3/ijrras_5_3_02.pdf

67
Zhang, L., Zhang, H., Ying, W., and Fang, D. 2013. Dehydration of methanol to dimethyl ether over Al2O3
catalyst: Intrinsic kinetics and effectiveness factor. The Canadian Journal of Chemical Engineering. [Online]. 91,
pp.1538-1546. [Accessed 24 June 2017]. Available from:
http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.221.9021&rep=rep1&type=pdf
Ziyang, B., Ma, H., Zhang, H., Ying, W., and Fang, D. 2013. Process simulation of dimethyl ether synthesis via
methanol vapour phase dehydration. Polish Journal of Chemical Technology. [Online]. 15(2), pp.122-127.
[Accessed 13 June 2017]. Available from: http://psjd.icm.edu.pl/psjd/element/bwmeta1.element.-psjd-doi-
10_2478_pjct-2013-0034

68
1

SYNGAS CONVERSION FROM BIOMAS REFORMING TO PRODUCE AN


ALTERNATIVE DIESEL FUEL FOR INDIA

Bibi Nariefa Abrahim and Valerie Dupont


School of Chemical and Process Engineering, University of Leeds, LS2 9JT, UK
Abstract: The kinetics of methanol synthesis and dehydration to produce dimethyl
ether using the Cu/ZnO/Al2O3 and -Al2O3 catalysts were studied in a catalytic
single-pass, shell and tube, fixed-bed reactor. The models proposed were based on
the hydrogenation of the carbon oxides, reverse water-gas shift reaction followed
by the dehydration of methanol using a two-staged production process. Sensitivity
analysis of temperature, pressure and feed-gas compositions were applied to each
model. A complete simulation of the production process was conducted to produce
catalyst and reactor specifications for each stage. The analysis of the models
indicated that feed-gas composition has a major impact on the product composition
and should therefore be a consideration for practical applications. The results
revealed that the processes are limited by thermodynamics, where the exothermicity
of the reactions result in favourable conversions at low temperatures. It was also
evident that the non-equimolar reactions were affected by pressure changes, where
higher pressures were preferred for greater methanol formation; unlike the
dehydration reaction, where the reaction kinetics excluded partial pressures and the
equimolar chemical reaction indicated that pressure variations had no influence on
the equilibrium conversions and dimethyl ether production. The models were found
to be in agreement with findings presented in literature, where minimal deviations
were noted, indicating that the kinetics employed in this study can be applied for
commercial production of dimethyl ether from biomass-derived syngas.
Keywords: Methanol synthesis, Methanol dehydration, Dimethyl ether, Reactor
model, Simulation study

1. INTRODUCTION

1.1 Introduction
The intense changes of the global temperatures and climate along with energy security has led to the
exploration of various alternative energy sources. India, with an emerging economy and increasing
population is one of the greatest consumer of energy, where the transport sector, specifically diesel-
powered vehicles, contributes significantly to the consumption. The demand for diesel fuel is expected
to increase in the future, given the increased presence of diesel-powered vehicles across the country. A
trend of increased diesel usage would contribute to heightened levels of greenhouse gases GHG),
resulting in issues of air pollution and climate change. The issues of global warming and energy security
triggered the exploration of indigenous resources for cleaner, renewable and sustainable energy
production. A very promising technology is the utilization of biomass to produce syngas and further
dimethyl ether (DME), a low polluting and economically favourable substitute for diesel.

DME, also referred to as methyl ether and dimethyl oxide is a colourless, flammable vapour or liquid.
The alternative fuel shares many similarities with diesel; however, DME has been reported to have a
better ignitability than diesel, given its high cetane number and low auto-ignition temperature (Albrecht,
2004). Additionally, as compared to diesel and biodiesel, DME contains greater amounts of oxygen
resulting in enhanced air-fuel mixing and a cleaner combustion. The application of biological waste for
energy harnessing, specifically biomass reforming to produce syngas and possibly DME, is one such
research area, which may result in improved access to diesel energy with reduced GHG emissions. This
2

study was focused on the examination of the necessity for diesel alternatives in India, specifically DME.
Moreover, this research aimed at assessing and developing an ideal gas reactor model for the conversion
of biomass-derived syngas to DME, as an alternative to diesel fuel for India.

1.2 Literature

DME can be produced from biomass-reformed syngas in a single phase or double phase synthesis
process, where syngas is converted to methanol in the presence of a catalyst, followed by catalytic
dehydration. Syngas contains varying amounts of hydrogen (H2), carbon monoxide (CO), carbon
dioxide (CO2), and small quantities of other gases (Albrecht, 2004). The production of DME using a
two-staged process of methanol synthesis and dehydration, has been studied extensively over the years.
The very first model for methanol synthesis with the Cu/ZnO/Al2O3 catalyst was presented by Leonov
et al in 1973. In this model, the effects of CO2 reactions were ignored and much focus was given to the
hydrogenation of CO, noting it to be the key source of carbon in the synthesis process (Bussche and
Froment, 1996). In 1982, Klier et al., proposed a new model which included the reaction of CO2,
however, the impact of the reaction on the kinetics were deemed minimal (Bussche and Froment, 1996).
Klier et al (Bussche and Froment, 1996) reported that the due to the high adsorption abilities of CO2,
the reaction rate of methanol synthesis and catalyst properties are greatly affected at varying partial
pressure ratios. Villa et al., (1985), also produced kinetic models whereby the hydrogenation of CO2
was excluded. In that study, low pressure methanol synthesis was conducted using the Cu/Zn/Al2O3
catalyst, where the hydrogenation of CO and the reverse water-gas shift reaction were noted to be the
main reaction pathways in the system. Contrary to the previous studies, Graaf et al., (1988) investigated
the effects of CO, CO2 and the water-gas shift reactions during low-pressure methanol synthesis over
the Cu/ZnO/Al2O3 catalyst. Graaf et al., (1988) reported that due to the occurrence of the water-gas shift
reaction, the reaction mechanism for methanol synthesis includes both the hydrogenation of CO and
CO2. In this study, Graaf et al., (1988) examined the specific mechanisms for the three reactions,
specifically hydrogenation of the carbon oxides and the water-gas shift reaction, where the rate
controlling steps were used to derive the kinetic equations.

Much like methanol synthesis, many kinetic models have also been proposed for the dehydration of
methanol to DME. One of the very early models, presented by Bercic and Jevec (1993), was based on
a plug flow reactor with the utilization of -Al2O3 catalyst to dehydrate methanol. According to Bercic
and Jevec (1993), the intra-particle mass transport was the rate controlling mechanism for this reaction.
Bercic and Jevec (1993) produced a model for the rate per catalyst particle using concentrations of the
reaction species and other reaction parameters. It was also reported that Bercic and Jevec (1993) altered
the rate equation by expressing the concentrations in the form of mole fractions, where they applied
apparent coefficients in the reaction equation. Further to this, Sierra et al., (2013) proposed a methanol
dehydration mechanism with the use of the -Al2O3 catalyst, where the kinetics were based on mass
conservation in an isothermal, plug-flow system. Sierra et al., (2013) also stated that the methanol
dehydration reaction is elementary and that the effect of water (H2O) in the reaction mechanism is
critical to the kinetics.

2. SIMULATION OF REACTOR MODELS

2.1 Methanol Synthesis

In this study, the kinetics for methanol synthesis was sourced from Graaf et al., (1988). The model was
computed based on various assumptions that have been proposed including:
 Single-pass shell and tube, fixed-bed catalytic reactor was selected.
 Plug flow model was used.
 Isothermal conditions were applied.
3

 Axial and radial dispersion were ignored.


 Ideal gas phase was applied.
 Effectiveness factor was assumed to be one.
The chemical reactions that were used to derive the kinetic models employed in this study are:
CO + 2H2 CH3OH (1)
CO2 + 3H2 CH3OH + H2O (2)
CO2 + H2 CO + H2O (3)
The kinetic models for the hydrogenation of CO and CO2 and reverse water-gas shift reaction, based
on the findings of Graaf et al., (1988):

 
 3 f CH3OH 
k1 K CO  f CO f H 2 
 
2
1 
 f H 2 2 K P1 
rCO   
  K  
      f 
1
1  K CO f CO  K CO2 f CO2 f H 2  2 H 2O
   K 1 2  H 2O   (4)
   H2  
 3 f f 
k2 K CO2  f CO2 f H 2 2  
CH 3OH H 2O

  
f H 2 3/2 K P 3 
rCO 2 
  K H 2O  
 
1  K CO f CO  K CO2 f CO2  f H 2 1/2  
 K H 1/2  2 
 f H 0
  2   (5)
 f H O fCO 
k3 K CO2  fCO2 f H 2  2 
rH 2O   KP2 
  K H2 0  
 
1  K CO f CO  K CO2 f CO2  f H 2 1/2  
 K H 1/2  H 2O 
 f
  2   (6)

where ki is the kinetic rate constant (molkg-1s-1), Ki is the adsorption equilibrium constant (bar-1), Kpi is
the equilibrium constant based on partial pressures (bar) , ƒi is the fugacity as partial pressures (bar).

The kinetic rate constant (k) and adsorption equilibrium constants are functions of temperature, where
the Arrhenius equation was used applied in the calculation as shown below:
𝐵 𝐵
ki = A exp (𝑅𝑇) Ki = A exp (𝑅𝑇) (7)

In both equations, R refers to the ideal gas constant (8.314 Jmol-1K-1) and T is the temperature (K). The
values of A and B are displayed in the table 2.1:

Table 2.1 Kinetic Parameters for Methanol Synthesis


Parameter A B
07
K1 2.69×10 -109900
K2 7.31×1008 -123400
02
K3 4.36×10 -65200
-07
KCO 7.99×10 58100
KCO2 1.02×10-07 67400
1/2 -11
KH2O/KH2 4.13×10 104500

The reaction equilibrium constants for the methanol synthesis were also sourced by Graaf et al.,
(1988), where the following equations were employed:
4

5139 2073
Log10 Kp1   12.621 Log10 Kp2   2.029 Kp3  Kp1  Kp2 (8)
T T
The kinetics were used to calculate the reaction rates along the reactor, resulting in the determination
of the flow rates, the rate of change of flow rates and the catalyst weights along the reactor.

2.2 Methanol Dehydration


The kinetics and reaction pathways employed for the dehydration of methanol to DME, were sourced
from Bercic and Jevec (1993). In the application of this model, similar assumptions as the methanol
synthesis process, with respect to reactor type, reaction conditions and mechanisms were applied. The
chemical equation considered in the simulation of the methanol dehydration model is shown below:

2CH3OH CH3OCH3 + H2O

The kinetic model for this chemical reaction that was selected is displayed below:

 yCH3OCH3 yH 2O 
kK 2CH3OH  yCH3OH 2  
 K eqm
r  

 
(9)
 
0.5 4
1  2 KCH3OH yCH3OH  K H 2O yH 20

where k is the reaction rate constant (molkg-1s-1), Ki is the adsorption equilibrium constant, Keqm is the
equilibrium constant and yi is the mole fraction of species ‘i’.
The kinetic constants expressed in the rate equation were computed using equations that were also
proposed by Bercic and Jevec (1993) and are presented in table 2.2:

Table 2.2 Kinetic Constants used in the Methanol Dehydration Model (10)
Kinetic Parameter Equation
Kinetic Rate Constant  10800 
k  6.60  108 exp  
 T 
ln  Keq   0.86 log T 
Equilibrium Constant 3138
 1.33 103 T  T 2  3.5 1010 T
T
Adsorption Equilibrium  830 
Constant K CH3OH  0.72  102 exp  
 T 
 1130 
K H 2O  0.45 102 exp  
 T 

2.3 Conversion Rates and Weights of Catalyst


The conversion of CO, CO2 and methanol were calculated based on the following relationship, where
Fi,in is the inlet molar flow of ‘i’ (mols-1) Fi,enditer is the molar flow rate at the end of iteration (mols-1):
𝐹𝑖,𝑖𝑛−𝐹𝑖,𝑒𝑛𝑑𝑖𝑡𝑒𝑟
Conversion (%)= 𝐹𝑖,𝑖𝑛
(11)

The weights of catalyst, W (kg) and the rates of change in weight of catalyst, dW (kg) were determined
using the following relationships:
5

Table 2.3 Determination of Catalyst Weights (12)

Methanol Synthesis 𝑑𝐹𝐶𝐻3𝑂𝐻


dW =
𝑟1+𝑟2
𝑑𝐹𝐶𝐻3𝑂𝐶𝐻3
W= dW + W
Methanol Dehydration dW= 𝑟

where dFCH3OH is rate of change in flow rate of methanol (mols-1), dFCH3OCH3 is the rate of change in
flow rate of DME (mols-1), r1 is rate of CO hydrogenation (molkg-1s-1), r2 is rate of CO2 hydrogenation
(molkg-1s-1) and r is rate of methanol dehydration (molkg-1s-1).

3. RESULTS AND DISCUSSION

3.1 Methanol Synthesis

The main findings of the methanol synthesis model are presented in table 3.1, where reactor, catalyst
and flow stream properties are shown:

Table 3.1 Main Characteristics of Methanol Synthesis Model

Property Unit Value


Inlet Temperature K 483
Pressure Bar 85
Number of Tubes ---- 3000
Reactor Length m 0.16
Reactor Diameter m 0.019
Catalyst Particle Diameter m 0.003
Catalyst Density kgm-3 1770
Catalyst Weight kg 0.08
Total Inlet Molar Flow Rate mols-1 0.64
Composition: CO/CO2/H2/H2O ---- 10.4/4.6/83.5/0.5

The system was modelled at a lab-scale, where the length and diameter of the reactor were found to
have an L/D ratio of 8.21, a value that confirms to the recommended rule of thumb for plug flow
reactors. It must be noted that a major influence on the model predictions and reaction rates is the
composition of the feed gas. In this scenario, the selected model is based on an inlet gas composition
that is similar to that produced through partially oxidised biomass-derived materials. The main findings
of this study are presented in the figures below:

Fig. 3.1 Reaction Species along Reactor Fig. 3.2 Reaction Rates along Reactor
6

In figure 3.1, it can be seen that CO was being consumed at a greater rate than CO2, indicating that a
CO-rich feed-gas maybe be favourable for methanol synthesis. Water production was also noticeable;
however, the levels were far less as compared to methanol, given that both hydrogenation reactions
resulted in the formation of methanol. In figure 3.2, rate 1 refers to the hydrogenation of CO, rate 2 as
hydrogenation of CO2 and rate 3 being the reverse water-gas shift reaction. It was evident that as the
hydrogenation reactions proceeded along the reactor, the rates at which they occurred decreased
gradually. The reverse water-shift gas reaction, however, produced a negative reaction rate, implying
that the reaction occurred in the opposite direction, where the products, CO and H2O were being
converted to CO2 and H2. This mechanism may have resulted in minimal contributions to the water
formation, much unlike carbon dioxide, which may have increased due to the rate of reverse water-gas
shift reaction. The enhanced CO2 levels may have resulted in increased hydrogenation of CO2, where
the reaction rate became greater than that of CO after some time.
Sensitivity analyses of the kinetic model were conducted with respect to pressure and temperature.

Fig. 3.3 Effects of Pressure and Temperature on the Formation of Methanol


Figure 3.3 clearly shows that at higher pressures, the formation of methanol in the reactor was greater,
a condition that that is thermodynamically controlled. In the hydrogenation reactions, the moles of the
reactants were greater than that of products resulting in a shift of equilibrium to the products as pressure
was increased. With respect to temperature, it was reported that the conversion rate of CO decreased as
temperature increased, indicating that the system favoured low-temperature environments. The
hydrogenation reaction of CO favoured a lower temperature for greater conversion, compared to that of
CO2, which initially increased with temperature up to a certain range and then decreased thereafter. The
sensitivity of the system to lowered temperature is connected to the exothermicity of the reactions,
where heat energy was an added product. In this study, methanol yields were found to be greater at
lower temperatures, given that the equilibrium constants decrease in high temperature environments.

3.2 Methanol Dehydration


The application of the kinetics for methanol dehydration resulted in the following findings:
Table 3.2 Properties of the Methanol Dehydration Model

Property Unit Value


Inlet Temperature K 573
Pressure Bar 2.1
Number of Tubes ---- 3000
Reactor Length m 0.48
Reactor Diameter m 0.056
Catalyst Particle Diameter m 0.003
Catalyst Density kgm-3 1130
Catalyst Weight kg 1.33
Total Inlet Molar Flow Rate mols-1 0.64
7

The reactor was modelled at a lab-scale where the length and diameter of the reactor were found to have
an L/D ratio of 8.54, a value that confirmed to the recommended rule of thumb for plug flow reactors.
The selected model is based on an inlet gas composition where CH3OH= 1, resulting from the previous
process of methanol synthesis. The reaction species along the reactor, followed by the reaction rate were
examined and is displayed below:

Fig. 3.4 Reaction Species along Reactor Fig. 3.5 Reaction Rate along Reactor
Figure 3.4 indicated that methanol concentration decreased as the reactor length increased, while the
levels of DME and water increased. Additionally, it can be seen that the production of DME is much
greater than that of water, given the selectivity of the Al2O3 catalyst. With respect to the reaction rate
and DME formation, it is visible that as the reactions were initiated, and the rates were higher, the
formation of DME rose steadily; however, as the rates slowed across the reactor, the DME level
followed closely, as the system approached equilibrium.
The sensitivity analysis revealed that pressure had no effect on the equilibrium conversion of methanol
and reaction rates, given the equimolar reaction and the exclusion of partial pressures in the rate
equation.

Figure 3.6: Effects of Temperature on the Formation of DME


The analysis of the temperature effects on the reactor model indicated that the conversion rate of
methanol to DME decreased as temperature increased, indicating that the system favoured low-
temperature environments. The favourable conversions at lower temperatures are due to the
exothermicity of the dehydration reaction, where the added heat reduces the equilibrium constant in
high temperature environments. However, it was reported that at very low temperatures, vast amounts
of catalyst are required to enhance the reaction, a situation which may be impractical for industrial
production of DME.
The models of methanol synthesis and dehydration were validated against literature where Jianqing et
al., (2002) and Dagde and Slyvanus (2016), indicated 1.2% and 1.3% deviations from the models,
8

respectively. The minimal deviations from this study with that presented in literature indicated that the
models are fairly accurate and can therefore be used in the simulation of DME production.
4. CONCLUSION
This study investigated the conversion of biomass-derived syngas to produce an alternative diesel fuel
for India. The simulation studies for the alternative fuel production were conducted in two stages:
methanol synthesis and methanol dehydration. In the methanol model, it was found that the
hydrogenation of CO occurred at the greatest rate as compared to the other reactions, and favoured
much lower temperatures than CO2. The rate of the reverse water-gas shift reaction was reported to be
negative, indicating that the reaction was moving in the opposite direction. The review of the synthesis
model also indicated that the system is thermodynamically controlled, with favourable conversions at
lower temperatures and higher pressures. The conditions of 483K and 85 bar produced the greatest
equilibrium production of methanol with appropriate catalyst and reactor requirements. The methanol
dehydration model reported that the production of DME was the greatest with increased methanol
composition in the feed-gas, however the catalyst and reactor requirements were slightly higher as
compared to other inlet gas ratios. The dehydration model was also found to be thermodynamically-
limited as greater conversions were reported at lower temperatures; however, very low temperatures
would require vast amounts of catalyst for an efficient production. Pressure changes to the dehydration
model were found to have no effect on the reaction rates and equilibrium conversions. The operating
conditions of 573K were found to be ideal for optimal DME production, given acceptable catalyst and
reactor requirements.

ACKNOWLEDGEMENT

The ESPRC is acknowledged for the GCRF grant EP/P51097X/1 (Integrated Low Carbon Energy
Solutions for Remote Rural Area Project).

REFERENCES

Albrecht, B.A. (2004). Reactor modelling and process analysis for partial oxidation of natural gas.
PhD Thesis, University of Twente.
Bercic, G., and J. Levec (1993). Catalytic dehydration of methanol to dimethyl ether. A kinetic
investigation and reactor simulation. IECR, 1993(32), pp.2478-2484.
Bussche, K.M.V., and G.F. Froment (1996). A steady-state kinetic model for methanol synthesis and
the water-gas shift reaction on a commercial Cu/ZnO/Al2O3 catalyst. J. Catal., 161(1996), pp.1-10.
Dadge, K.K. and H.U. Slyvanus (2016). Modelling of a tubular fixed-bed reactor for the production of
dimethyl ether using alumina catalyst. IJCPER, 3(2), pp.23-34.
Graaf, G.H., E.J., Stamhuis, A.C.C.M., Beenackers (1988). Kinetics of low-pressure methanol
synthesis. Chem. Eng. Sci., 43(12), pp.3185-3195.
Jianqing, Z., T., Noritatsu, J., Kaoru (2002). New process of low-temperature methanol synthesis
from CO/CO2/H2 based on dual-catalysis. Sci. China, 45(1), pp. 106-112.
Sierra, I., J. Erena, A.T. Aguayo, A. Ateka and J. Bilbao (2013). Kinetic modelling of dehydration of
methanol to dimethyl ether over y-Al2O3 catalyst. Chem. Eng. Trans, 32, pp.613-618.
Villa, P., P. Forzatti, G. Buzzi-Ferraras, G. Garone and I. Pasquon (1985). Synthesis of alcohols from
carbon oxides and hydrogen 1. Kinetics of the low-pressure methanol synthesis. Ind. Eng. Chem.
Process Dev., 24(1), pp.12-19.

View publication stats

You might also like