Santi Kulprathipanja, James E. Rekoske, Daniel Wei, Robert v. Slone, Trung Pham, Chunqing Liu - Modern Petrochemical Technology - Methods, Manufacturing and Applications-Wiley-VCH (2021)
Santi Kulprathipanja, James E. Rekoske, Daniel Wei, Robert v. Slone, Trung Pham, Chunqing Liu - Modern Petrochemical Technology - Methods, Manufacturing and Applications-Wiley-VCH (2021)
Santi Kulprathipanja, James E. Rekoske, Daniel Wei, Robert v. Slone, Trung Pham, Chunqing Liu - Modern Petrochemical Technology - Methods, Manufacturing and Applications-Wiley-VCH (2021)
Santi Kulprathipanja
James E. Rekoske
Daniel Wei
Robert V. Slone
Trung Pham
Chunqing Liu
Authors All books published by WILEY-VCH
are carefully produced. Nevertheless,
Dr. Santi Kulprathipanja authors, editors, and publisher do not
Honeywell UOP warrant the information contained in
25 East Algonquin Road these books, including this book, to
IL be free of errors. Readers are advised
United States to keep in mind that statements, data,
illustrations, procedural details or other
Dr. James E. Rekoske items may inadvertently be inaccurate.
Ecolab Inc.
1 Ecolab Place Library of Congress Card No.:
St. Paul, Minnesota 55102 applied for
United States
British Library Cataloguing-in-Publication
Dr. Daniel Wei Data
Honeywell UOP A catalogue record for this book is
25 East Algonquin Road available from the British Library.
IL
United States Bibliographic information published by
the Deutsche Nationalbibliothek
Dr. Robert V. Slone The Deutsche Nationalbibliothek lists
Honeywell Advanced Mateirals this publication in the Deutsche
25 East Algonquin Road Nationalbibliografie; detailed
IL bibliographic data are available on the
United States Internet at <http://dnb.d-nb.de>.
10 9 8 7 6 5 4 3 2 1
v
Contents
Foreword xi
Preface xiii
Part I Introduction 1
2 Petrochemical Markets 17
2.1 Introduction 17
2.2 The Market for Ethylene 18
2.3 The Market for Propylene 22
2.4 The Market for C4 Olefins and Diolefins 26
2.5 The Market for Aromatic Compounds 32
2.5.1 The Market for Benzene 32
2.5.2 The Market for Toluene 34
2.5.3 The Market for Xylene 36
2.6 The Market for Synthesis Gas 37
vi Contents
7 Adsorption 147
7.1 Principle of Adsorption 147
7.2 Adsorbents 148
7.2.1 Zeolite Adsorbents 148
7.2.2 Zeolite Synthesis 150
7.2.3 Forming Zeolite Powders into Usable Shapes 152
7.2.3.1 Pelletization of Zeolite Powder to Extrudate Adsorbent 153
7.2.3.2 Accretion of Zeolite Powder to Bead Adsorbent 153
7.3 Bulk Liquid-Adsorptive Separation 154
7.3.1 Modes of Liquid-Adsorptive Separation 154
7.3.1.1 Adsorption Isotherms 156
7.3.1.2 Pulse-Test Procedure 156
7.3.1.3 Breakthrough Procedure 157
7.3.1.4 Simulated Moving-Bed Operation: Sorbex Technology 158
7.3.2 Liquid-Adsorptive Separation Processes 159
7.3.2.1 Equilibrium-Selective Adsorption 159
7.3.2.2 Rate-Selective Adsorption 166
7.3.2.3 Shape-Selective Adsorption 168
7.3.2.4 Ion Exchange 169
7.3.2.5 Reactive Adsorption 170
7.4 Commercial Bulk Liquid-Adsorptive Process 170
7.4.1 Parex 170
7.4.2 MX SorbexTM 173
Acknowledgment 174
References 174
8 Distillation 179
8.1 Introduction to Distillation 179
8.2 Principles and Systems 181
8.2.1 Vapor–Liquid Equilibrium (VLE) Theory 181
8.2.2 McCabe–Thiele Diagram 182
8.2.3 Factors for VLE-Based Distillation Design 184
8.2.3.1 Relative Volatility 184
8.2.3.2 Activity Coefficient 185
8.2.3.3 Solubility 185
8.2.3.4 Azeotropes: Maximum Concentration Limit 185
8.2.3.5 Surface Area 186
Contents ix
9 Membranes 203
9.1 Principle and Background 203
9.2 Types and Preparation of Membranes 204
9.2.1 Polymeric Membranes 205
9.2.1.1 Fundamentals of Polymeric Membranes 205
9.2.1.2 Preparation of Polymeric Membranes 207
9.2.2 Inorganic Membranes 213
9.2.2.1 Introduction to Microporous Crystalline Inorganic Membranes 214
9.2.2.2 Preparation of Microporous Crystalline Inorganic Membranes 215
9.2.3 Mixed-matrix Membranes 218
9.2.3.1 Introduction to Mixed-matrix Membranes 219
9.2.3.2 Preparation of Mixed-matrix Membranes 221
9.3 Membrane Modules 223
9.4 Membrane Reactors 225
9.5 Commercial Applications of Membranes 227
References 235
10 Absorption 241
10.1 Introduction to Absorption 241
10.2 Acid Gas Removal 243
10.3 Chemical Solvent 245
10.3.1 Amines Process 246
10.3.1.1 Monoethanolamine (Primary Amine) 248
10.3.1.2 Diethanolamine (Secondary Amine) 248
10.3.1.3 Methyldiethanolamine (Tertiary Amine) 248
10.3.2 BenfieldTM Process 249
10.4 Physical Solvent 251
10.4.1 Solvent Selection 252
x Contents
Index 283
xi
Foreword
Over the past 30 years, there has been a dramatic global expansion of the petro-
chemicals industry. The industrialization and economic growth of countries in
Asia, the Middle East, Latin America, and now Africa has raised the purchasing
power of billions of people and created surging demand for products that are
ultimately derived from petrochemical feedstocks. As a supplier of technology to
the petrochemicals industry, Honeywell UOP has been a willing participant in this
expansion and has worked with its customers around the world to deploy safe and
state-of-the-art processes for making the basic monomers and chemicals that go
into making the thousands of products that we all consume every day.
Dr. Santi Kulprathipanja has been at the heart of developing many of Honeywell
UOP’s technologies. In his very distinguished career, Dr. Kulprathipanja was named
as inventor of over a hundred patents, including many of the technologies described
in this book. While best known for his many inventions, Dr. Kulprathipanja was
always a strong advocate for developing close links between industry and academia,
and for many years he has also taught university classes around the world so that he
could share his knowledge with the next generation of engineers and chemists.
This book builds on Dr. Kulprathipanja’s highly regarded classes in petrochemicals
processing and separation technologies and contains additional contributions from
colleagues at Honeywell UOP who have joined him in his teaching and provided
additional insights from their own technical expertise. It should be a very useful
reference for chemical engineers or others who might be starting a career in the
petrochemicals industry, or in companies or institutions that serve that industry.
Gavin Towler
Chief Technology Officer and
Vice President of Research and
Development Honeywell UOP
xiii
Preface
This book will provide an overview of markets and uses for common petrochemical
building blocks and also a survey of the technology used to make the key petrochemi-
cal building blocks. The book will be written with a practical chemistry and chemical
engineering course work in mind, with the target audience being university profes-
sors, students, scientists, and engineers those that are in the field of petrochemicals
and petrochemical technology. The book will contain both critical information to
help the readers to understand how the technologies are used to produce petro-
chemicals, how the various petrochemicals are used, and examples and problems
designed to reinforce the learning about the basic science, engineering, and use of
petrochemicals.
The topics covered by the book include Chapter 1: Refining and Petrochemical
Process, Chapter 2: Petrochemical Markets, Chapter 3: Olefins and Synthesis Gas
Production Technologies, Chapter 4: Uses of Olefins and Synthesis Gas, Chapter 5:
Aromatic Production Technology, Chapter 6: Uses of Aromatics, Chapter 7: Adsorp-
tion, Chapter 8: Distillation, Chapter 9: Membranes, Chapter 10: Absorption, and
Chapter 11: Extraction. Highlights in each chapter can be found as given in the fol-
lowing content.
Refining and petrochemical process is introduced in Chapters 1 and 2. As we can
notice in the last 10 years or so, the combination of high demand for electric cars
and higher automobile engine efficiency in the future will mean less conversion of
petroleum into fuels. However, the demand for petrochemicals is forecast to rise due
to the increase in world population. With this, it is expected that modern and more
innovative technologies will be developed to serve the growth of the petrochemi-
cal market. Petroleum is refined/converted into intermediate refinery products. The
intermediate refinery products are then converted to fuels such as fuel gas, liquefied
petroleum gas, gasoline, jet fuel, kerosene, auto diesel, and other heavy products. In
addition, these petroleum intermediates can be further processed and separated into
products for petrochemical applications.
We start our journey toward a better understanding of petrochemical markets
through an exploration of the building blocks of the industry – those commodity
chemicals that are utilized in large volume – in Chapter 1. We introduce the
sources of these building blocks – predominantly from three separate but important
complexes: olefin complexes, aromatic complexes, and synthesis gas complexes.
xiv Preface
We also introduce the “interconnected” nature of these building blocks – the fact
that their sources and uses are frequently connected in which one building block is
co-produced with another. In Chapter 2, we begin to explore this interconnection in
more detail, providing some statistics for each building block. Many of the building
blocks have many different sources, are made as co-products with other building
blocks and chemicals, and the supply of these building blocks depends rarely just
on one of the building blocks. Because of this, the markets frequently get out of
balance – where either shortages or large surpluses of one building block are noted.
The beauty of these markets then comes through, as substitution and movements
between the building blocks occur to rebalance the supply and demand. The ability
to make many final products (different products) from one building block, as well as
the ability to use many building blocks to make one final product (different routes)
make the petrochemical markets extremely complex and extremely robust.
Olefins are key building blocks of chemical industry. Light olefins, ethylene and
propylene, are the key components used in the manufacture of many polymers
and chemicals. Butylene is also in high demand for further processing to pro-
duce methyl-tert-butyl ether (MTBE) and alkylate for gasoline blending. Heavy
normal-olefins (C10 –C14 ) are the main feedstock component of linear alkylbenzene
(LAB), the most common raw material in the manufacture of biodegradable
household detergents.
Production and uses of olefins are documented in Chapters 3 and 4. They
include the olefins productions using steam cracking, paraffin dehydrogenation,
and methanol to olefins technologies as well as the uses of the olefins to produce
downstream products such as polyolefins, polyvinyl chloride, mono-ethylene glycol,
propylene oxide, acrylonitrile, LAB, MTBE, and alkylate. Chapters 3 and 4 also
cover the manufacturing technologies for producing syngas, which is used in large
volumes for the production of commodity chemicals such as hydrogen, ammonia,
and methanol.
Aromatics are important building blocks for petrochemical intermediate and
derivative products. Chapters 5 and 6 describe various methods and commercial
processes to produce aromatics such as benzene, toluene, and xylene (BTX).
Naphtha reforming is still the most efficient and sensible method to produce aro-
matics since direct dehydrogenation of naphthenes is favorable provided sufficient
temperature while valuable hydrogen co-product can be obtained at the same time.
Other catalytic processes include fluidified catalytic cracking, aromatization of
light naphtha, toluene methylation, aromatic isomerization, transalkylation, and
disproportionation. Noncatalytic methods include steam cracking, pyrolysis, and
coke-oven benzole production.
Essential materials such as nylons, polyesters, plastics, pesticides, coatings, and
resins are primarily derived from basic aromatic building blocks, including BTX.
These materials have strong mechanical strength and thermal stability with aro-
matic ring in the backbone. Electrophilic aromatic substitution allows functional
groups to attach to the aromatic molecules that make them flexible monomers for
further production or synthesis of polymers for construction and automotive indus-
tries or consumer and pharmaceutical products.
Preface xv
Membrane technology has been a critical enabler for effective and efficient gas,
vapor, and liquid separations for many years. The membrane technology is described
in Chapter 9. Polymeric membrane–based industrial separation processes have
evolved rapidly since the development of asymmetric integrally skinned polymeric
membranes by Loeb and Sourirajan in 1960s and the development of thin-film
composite (TFC) membranes by Cadotte in 1980s. Inorganic membranes have the
advantages of high thermal stability, high chemical resistance, and resistance to
harsh environments over polymeric membranes. Therefore, inorganic membranes
not only have been commercially successful in some liquid separation applications,
but also have been studied for catalytic membrane reactor, energy storage, and
other applications. Further advancement in reducing the cost and improving the
processability and reproducibility of the inorganic membranes is still critical for
inorganic membranes to be adapted for new commercial applications. Innovation
in membrane materials with long-term stability, chemical resistance and specific
to applications, membrane fabrication methods, module design, new membrane
process design, and process technology integration has been instrumental to the
widespread industrial adoption of membrane technology, and it will continue to
enable new membrane-based separations and reactions. Chapter 9 will discuss sev-
eral aspects of membranes, including principles, types and preparation, membrane
modules, membrane reactors, and commercial applications.
Absorption technology is given in Chapter 10 to detail method of separation com-
monly used in the oil and gas and chemical industries for removal of acid contami-
nants such as H2 S (hydrogen sulfide) and CO2 (carbon dioxide) from hydrocarbons;
this separation strategy is often employed in the purification, transport, and storage
of natural gas. As natural gas usage has increased over the past 20 years due to its
lower CO2 emissions vs. coal, the removal of acid gases has become more critical
and helps protect large capital investments from corrosion.
There are two primary classes of absorption we cover in Chapter 10: chemical
absorption and physical absorption. The strengths and weaknesses of each method
are discussed. At low partial pressures of the component (generally low concentra-
tion), a chemical solvent is generally preferred while at high partial pressures of
the component to be removed, physical solvent has advantage due to its high load-
ing capability. The criteria by which a specific amine is selected are reviewed: (i)
Required purity of cleaned gas stream for further processing or end use, (ii) Com-
position of feed gas, (iii) Utility requirements, process costs, and (iv) Corrosion and
solvent degradation/solvent losses.
We then turn our attention to CO2 removal using an inorganic chemical solvent
(K2 CO3 in the UOP BenfieldTM process). The Honeywell UOP Benfield Process is
a thermally regenerated cyclical solvent process using an activated, inhibited hot
potassium carbonate solution to remove CO2 , H2 S, and other acid gas components.
An overview of the physical solvent options is included: SelexolTM (DMPEG,
Dimethyl Ether of Polyethylene Glycol), RectisolTM (Methanol), and Fluor (PC).
Cost and performance factors such as solvent volatility, contaminant solubility,
and process conditions such as pressure and temperature often guide the choice of
solvent in physical absorption systems.
Preface xvii
Section I
Introduction
3
1.1 Introduction
The combination of high demand for electric cars and higher automobile engine effi-
ciency in the future will mean less conversion of petroleum into fuels. However, the
demand for petrochemicals is forecast to rise due to the increase in world popula-
tion. With this, it is expected that modern and more innovative technologies will be
developed to serve the growth of the petrochemical market.
In a refinery process, petroleum is converted into petroleum intermediate prod-
ucts, including gases, light/heavy naphtha, kerosene, diesel, light gas oil, heavy gas
oil, and residue. From these intermediate refinery product streams, several fuels
such as fuel gas, liquefied petroleum gas, gasoline, jet fuel, kerosene, auto diesel, and
other heavy products such as lubricants, bunker oil, asphalt, and coke are obtained.
In addition, these petroleum intermediates can be further processed and separated
into products for petrochemical applications.
In this chapter, petroleum will be introduced first. Petrochemicals will be intro-
duced in the second part of the chapter. Petrochemicals – the main subject of this
book – will address three major areas, (i) the production of the seven cornerstone
petrochemicals: methane and synthesis gas, ethylene, propylene, butene, benzene,
toluene, and xylenes; (ii) the uses of the seven cornerstone petrochemicals, and (iii)
the technology to separate petrochemicals into individual components.
1.2 Petroleum
Petroleum is derived from the Latin words “petra” and “oleum,” which means “rock”
and “oil,” respectively. Petroleum also is known as crude oil or fossil fuel. It is a
thick, flammable, yellow-to-black mixture of gaseous, liquid, and solid hydrocarbons
formed from the remains of plants and animals. Over millions of years, this organic
mixture of plants and animals was subjected to enormous pressure and heat as layers
of earth further buried them over time. The organic mixture changed chemically and
broke down into hydrocarbons. Because of the time it takes to form petroleum, it is
referred to as a nonrenewable energy source.
1.2.1.2 Liquid Petroleum: Liquid Petroleum Can Be Separated into Light Crude
Oil and Heavy Crude Oil
Light crude oil has a low viscosity and is a liquid at room temperature. These prop-
erties make light crude oil easy to pump and extract. It is composed of short-chain
paraffins, which are straight and branched-chain hydrocarbons. Because virgin light
crude oil comprises these short chains, it does not have to be heavily refined to pro-
duce gasoline. Typically, the chain-length range in gasoline is 4–12 carbons, making
light oil a desirable crude to be processed in a refinery. Approximately 30% of the
world’s petroleum reserves is light oil crude.
Heavy crude has a higher viscosity, but is still a liquid at room temperature. It usu-
ally contains more sulfur, nitrogen, and other contaminants than light oil. To refine
heavy oil to gasoline, it must be cracked and treated to remove the contaminants.
It requires more energy input and cost. Beyond the need for additional refinement,
heavy crude also needs additional extraction techniques to recover oil from the wells.
These techniques include stream stimulation to make the oil less viscous and the
injection of air into the wells to create fires that burn heavier hydrocarbons and
degrade them into lighter, more easily pumped compositions.
The transport of heavy oil requires the addition of diluting agents, particularly
in pipelines. The other major drawback to heavy crude is its environmental impact
because it contains sulfur and heavy metals, both of which must be removed.
1.2 Petroleum 5
Heavy metals are often toxic and their removal from crude presents disposal issues.
Sulfur, which may be as high as 4.5 wt%, is corrosive to pipeline metal and refinery
components.
Element Percent
Carbon 83–85
Hydrogen 10–14
Nitrogen 0.1–2.0
Oxygen 0.05–1.5
Sulfur 0.05–6.0
Metals <0.1
Hydrocarbon Percent
Alkanes 15–60
Naphthenes 30–60
Aromatics 3–30
Asphaltenes Remaining
processing units and auxiliary facilities such as utility units and storage tanks. Some
modern petroleum refineries process as much as 900 000 barrels per day of crude
oil. According to Oil and Gas Journal, a total of 636 refineries operated worldwide
at the end of 2014 for a total capacity of 87.75 million barrels per day of crude oil.
Reliance Industries’ Jamnagar Refinery in Gujarat, India, currently is the world’s
largest oil refinery. Each refinery has its own unique arrangement and combination
of refining processes largely determined by the refinery location, desired products,
and economic considerations.
Conversion of heavy oils to useful products requires breaking large molecules into
smaller ones. The breaking or cracking of large molecules can be accomplished by
one or more combinations of heat, pressure, and chemical reaction. In the process,
harmful or unwanted compounds such as metals, sulfur, nitrogen, and oxygen are
also removed. The conversion of heavy oils into useful products requires many sep-
aration and chemical reaction processes. These processes are briefly characterized
as follows:
● Desalination: This is the first unit in the refinery complex. Salts from crude oil are
extracted before it enters the atmospheric distillation unit.
● Crude oil distillation (atmospheric distillation): The desalted crude oil is separated
into various fractions for processing in downstream units.
● Vacuum distillation: The residue oil from the bottom of the crude oil distillation
unit is further distilled at a vacuum pressure well below atmospheric.
● Naphtha hydrotreater: The hydrotreater desulfurizes, denitrogenizes, and deoxyg-
enizes naphtha using hydrogen from the atmospheric distillation unit. The naph-
tha must be treated before sending the stream to the catalytic reformer unit.
● Catalytic reformer: The reformer converts the hydrotreated naphtha into refor-
mate, which has a higher content of aromatics and cyclic hydrocarbons. The end
products are high-octane gasoline and para-xylene aromatics, which is a critical
petrochemical in making PET.
● Distillate hydrotreater: This unit, similar to the naphtha hydrotreater, removes sul-
fur, nitrogen, and oxygen from distillates (such as diesel) and other units within
the refinery after atmospheric distillation.
● Fluid catalytic cracker (FCC): The FCC process is used to upgrade the heavier,
higher boiling-point fractions from the crude oil distillation by converting them
into more valuable lighter and lower boiling-point products.
● Hydrocracker: The hydrocracker uses hydrogen to upgrade heavy residual oils
from the vacuum distillation unit by thermally cracking them into more valuable
lighter and lower-viscosity products.
● Merox: The Merox process desulfurizes LPG, kerosene, and jet fuel by oxidizing
mercaptans to organic disulfides.
● Coking: Delayed cokers, fluid cokers, and flexicokers are used to crack very heavy
residual oils into gasoline and diesel fuel leaving petroleum coke as a residual
product.
● Alkylation: Alkylation is a process that uses sulfuric or hydrofluoric acid or an
ionic liquid as a catalyst to produce high-octane components for gasoline blending.
8 1 Refinery and Petrochemical Processes
Liquefied
petroleum gas
Light straight-run naphtha
Normal paraffins
Desalter
Asphalt Asphalt
oxidation
For example, iso-butane and butylenes can be converted into iso-paraffin alkylate,
which is a very high-octane gasoline.
● Dimerization: Dimerization is used to convert olefins into higher-octane gasoline.
For example, butenes are dimerized into isooctene, which is subsequently hydro-
genated to form isooctane.
● Isomerization: This process converts linear molecules such as normal pentane to
higher-octane branched molecules for blending into gasoline. Isomerization units
also are used to convert linear molecules such as normal butane into isobutane for
use in alkylation units.
Oil refining is the process where crude oil is transformed into more desirable and
valuable products using the processes described above. Refineries often are classified
by the number and type of process units available for transforming crude oil into
petroleum products.
Topping refineries are the least complex refineries and are used to separate the
crude oil into its constituent petroleum products by atmospheric distillation; naph-
tha is produced, but no gasoline. Hydroskimming refineries are also one of the sim-
plest types of refineries used in the petroleum industry and are equipped with atmo-
spheric distillation, naphtha reforming, and additional treating processes to produce
gasoline. Important to note, a hydroskimming refinery produces a surplus of fuel
with a relatively unattractive price and demand. Figure 1.1 shows schematic topping
and hydroskimming refineries.
The next refinery classification in complexity level is a medium-conversion
refinery. A medium-conversion refinery is shown in Figure 1.2. To design
a medium-conversion refinery, unit operations such as vacuum distillation,
thermocracking, fluid catalytic cracking, and asphalt oxidation are added to the
hydroskimming refinery. This added level of complexity allows conversion of fuel
oil to light distillates and middle distillates.
1.2 Petroleum 9
Atmospheric Catalytic
condensation
gas oil
Fluid catalytic Unsaturated Sulfur
cracking unit vapor recovery removal
Alkylation
Visbreaking Sulfur
thermal cracking removal C4 isomerization
Desalter MTBE
Vacuum
distillation
Crude oil column
Residual fuel oil
Atmospheric
residue
Asphalt Asphalt
oxidation
Hydrocracking Catalytic
Atmospheric unit condensation
gas oil
Fluid catalytic Unsaturated Sulfur
cracking unit vapor recovery removal
Alkylation
Visbreaking Sulfur
thermal cracking removal C4 isomerization
Desalter MTBE
Coking
Vacuum
distillation
Crude oil column
Demetallizing Residual fuel oil
Atmospheric
residue Coke
Dewaxing Lubricant Lubricants
compounding
Asphalt Asphalt
oxidation
The most complex type of refinery is the high-conversion refinery, which is shown
in Figure 1.3. The high-conversion refinery adds more inter-related processes such
as hydrocracking, coking, demetallizing, and dewaxing. In particular, a coking unit
adds cracking capability for conversion of fuel oil into distillates and petroleum coke,
allowing high-efficiency conversion of the crude oil feedstock into higher yields of
more valuable products.
Gasoline 46
Jet fuel 9
Diesel and other fuel 26
Heavy fuel oil 4
Lubricants 1
Asphalt 3
Other products 11
derived from petroleum refineries. Depending on the composition of the crude oil
and the demands of the market, refineries can produce different types and pro-
portions of petroleum products. Petroleum products are usually grouped into four
categories: light distillates (LPG, gasoline, naphtha), middle distillates (kerosene,
jet fuel, diesel), heavy distillates, and residuum (heavy fuel oil, lubricating oils, wax,
asphalt). The largest share of petroleum products is fuels, which include single
or blended combinations of hydrocarbons to provide gasoline, jet fuel, diesel fuel,
heating oil, and heavier fuel oils. Heavier fractions also can be used to produce
asphalt, tar, paraffin wax, lubricating, and other heavy oils. Table 1.3 summarizes a
breakdown of the petroleum products made from a typical barrel of U.S. oil.
Refineries also produce chemicals that are used to produce polymers, detergents,
and other useful consumer products. Since petroleum contains sulfur-containing
molecules, elemental sulfur is often produced as a petroleum byproduct. Carbon,
in the form of petroleum coke, and hydrogen gas also are produced as petroleum
byproducts. For example, hydrogen often is used as an essential reactant for other
oil refinery processes such as hydrocracking and hydrodesulfurization.
Chemical structures of these basic building blocks are shown in Figure 1.4.
From these seven building blocks, and adding abundant, readily available chemi-
cals such as air, pure oxygen, or pure nitrogen, the petrochemical industry has built a
wide variety of value-added compounds that have become essential to modern soci-
ety. For example, in a modern automobile, plastics and petrochemical compounds
H
H H3C CH3
H H H C CH3 CH3
C C H
C C CH2 H3C
H H
H H
H3C CH3
CH2
Ethylene Propylene 1-Butene 2-Butene Isobutene
CH3
CH2
H2C H2C C
H2 CO
1,3-Butadiene 1,2-Butadiene
Synthesis gas
C4 mono- and di-olefins
CH3
CH3 CH3
CH3
H3C CH3
CH3
Xylene isomers
Aromatic hydrocarbons
account for more than 50% of the volume of materials used. Despite this large frac-
tion of the volume, the petrochemical parts make up only about 10% of the vehicle
weight [7]. Clearly the trend of replacing automotive parts with lighter-weight pieces
based on petrochemicals will have to continue as lighter, more energy-efficient, and
electric vehicles become more popular.
The seven building blocks are mostly present as components of raw materials
(petroleum, natural gas) described earlier. Indeed, most of the building blocks had
been identified in the 1800s as components of these raw materials. For example,
Faraday [8] was reportedly the first to successfully isolate and characterize benzene
in 1825. However, these compounds were needed to be present in much larger quan-
tities, and isolated as intermediates for subsequent chemical reactions, for them to be
of use as building blocks. With petroleum and natural gas as starting materials, the
industry began to invent methods to obtain these building blocks in vast quantities.
Today, the unit operations (reactions, separations, heat transfer, etc.), which are
combined in systematic ways to make petrochemical building blocks in large quanti-
ties, often are referred to as petrochemical complexes. Although very few petrochem-
ical complexes are exactly alike, we can categorize these complexes into three basic
types.
● Olefin complexes: The complexes make predominantly the olefins ethylene, propy-
lene, and butenes. It is the primary technology for the product of olefins in steam
cracking [9], but other technologies such as dehydrogenation and methanol to
olefins technologies [10, 11] are gaining traction. In a steam cracking–based olefin
complex, the primary product is ethylene; the other products are formed in far
lower quantities. These complexes will sometimes be “fully integrated” – meaning
they have full recovery of propylene, butenes, and even the aromatics formed. Oth-
ers will recover only the ethylene and propylene and sell the remainder of the
liquid products to others for recovery. A sample block-flow diagram for an olefins
complex using steam cracking of hydrocarbons is shown in Figure 1.5.
● Aromatics complexes: These complexes make mostly aromatics hydrocarbons, and
often make most of the product as para-xylene for use in the synthesis of puri-
fied terephthalic acid, PTA [12]. These complexes usually employ a technology to
reform the naphtha cut of petroleum into aromatic rings, and then use a series
of technologies to interconvert the aromatic rings into the desired products, with
Ethane recycle
Methane
Propane recycle
C4 products
Debutanizer (to separation or sales)
Pyrolysis gasoline
(to aromatics extraction)
Raffinate
Sulfolane Benzene
extraction
H2 LE Toluene
C9 aromatics
H2 LE H2 LE
THDA
Naphtha CCR H2 LE
RS Clay BC TC HA
HDT platform
Tatoray
C10 +
ortho-Xylene
H2 LE Light ends
C7A
para-xylene being typically the most desired. A typical plant-flow diagram for an
aromatics complex is shown in Figure 1.6.
● Synthesis gas complexes: In these complexes, coal, natural gas, or some other
hydrocarbon source is steam reformed into a synthesis gas mixture (H2 and CO).
This synthesis gas then is converted into a variety of different products, including
methanol or ammonia by reacting the hydrogen with nitrogen recovered from
air. While methanol is a very-large-volume commodity chemical, ammonia
production is even larger due to the use of ammonia in agricultural fertilizers.
While the types of reformers and the layout vary significantly based on feedstock,
product ratios, and reformer technology, a typical layout for a synthesis gas plant
making methanol from natural gas is shown in Figure 1.7.
Together, these three types of complexes form most of the conversion of raw materi-
als into the building blocks, which support the petrochemical industry.
Many authors have tried to point to the “birth” of petrochemicals – something
which has always resulted in controversy. This is because petrochemicals have
existed for as long as our natural resources have existed, but were often masked
Water
Water Steam
pretreatment generation Purge gas (fuel)
Water
(to waste water treatment)
MTBE LDPE,HDPE,LLDPE
Methane Formaldehyde
Chlorine Poly(vinyl
EDC Vinyl chloride chloride)
Ethane Ethylene oxide Ethylene
Ethylene glycol Polyester
alpha-Olefins
Steam Propylene
Olefins
Propane cracking Ethanol
Polyisobutylene
Butylenes Isobutane
Polypropylene
Butanes Propylene oxide Polyols
Butadiene
Methyl methacrylate
n-Butane
Maleic
anthydride
SB latex and rubber
Alkyd resins
ortho-Xylene Phthalic anhydride Plasticizers
meta-Xylene Isophthalic acid Unsaturated
Xylenes
polyester resins
para-Xylene Purified terephthalic acid
Polyester
Dimethyl terephthalate fibers and resins
Chemicals
Ethylene dichloride Ethylene oxide
Vinyl chloride Ethylene glycol
Ethylbenzene Vinyl acetate
Styrene Ethanol
H H Oligomers α -Olefins
C=C Fatty alcohols
H H
Ethylene
References
Petrochemical Markets
2.1 Introduction
Describing the web of production and uses of petrochemicals often evokes a number
of different adjectives: complex, integrated, confusing are among the most common
descriptors. In our experience, however, through the introduction and explanation
of a few simple economic and technical concepts, the business end of petrochemical
production and use can be quite simple and exciting. The key to unpacking this com-
plexity is the concept of the core building blocks for modern petrochemicals, which
we introduced in Chapter 1. As a reminder, these core building blocks are:
● Ethylene
● Propylene
● C4 Mono- and Di-Olefins
● Benzene
● Toluene
● Xylenes
● Synthesis Gas
We call them the building blocks of the modern petrochemical industry because,
quite simply, nearly 90% of the volume of petrochemical products produced and
traded are derived from these seven compounds or mixtures [1].
The building blocks are often referred to as primary or commodity chemicals. Sim-
ply put, a commodity chemical is often described as a chemical material made at
large scale and which is used and traded globally. While it is true that the build-
ing blocks are, indeed, commodities, they are not the only commodity chemicals. So
why does this definition matter? Because commodities have commonalities, which
make them easy to trade – to sell between companies, and even between regions. For
commodity chemicals, these commonalities are often chemical specifications – what
concentration, what level of impurities, and what form (liquid, solid, solution, etc.)
are the typical specifications used in the commodity chemical industry. With this set
of properties to define the commodity, one company can buy the commodity from
one or more different suppliers at different times and have confidence that they can
use that commodity to make the end products they want to make.
Modern Petrochemical Technology: Methods, Manufacturing and Applications, First Edition.
Santi Kulprathipanja, James E. Rekoske, Daniel Wei, Robert V. Slone, Trung Pham, and Chunqing Liu.
© 2021 WILEY-VCH GmbH. Published 2021 by WILEY-VCH GmbH.
18 2 Petrochemical Markets
In Chapter 1, we also briefly described how these building blocks are made. In
addition to how they are made, how much and where (geographically) the building
block is made is also important to understand. Collectively, this information makes
up what can be called sources of supply, or just source or supply in short. Of course,
there is no use making a chemical or building block if there is no use for it, or demand
for the chemical. Sometimes this use is also called a sink, where the term “sink”
represents a use or consumption of the chemical in question. The spatial and time
balance and trade of these chemical commodities is done through markets. While
often not physical locations like they historically have been, markets are the pro-
cesses and institutions through which goods change hands. Matching sellers (sup-
ply) with buyers (demand) is the key function of the market. Often times, today, this
is done electronically without the need for the buyer and seller to meet or, in some
cases, even know each other. Through this process, the commodities, which have
common specifications, can be traded between companies, regions, or even globally.
While not the intent of this chapter, it is important that the reader understand
some basic fundamentals regarding the economic principle regarding the balance of
supply and demand [2]. Briefly, the process by which supply and demand are bal-
anced is through the market-clearing price [3]. Generally speaking, suppliers will
want to supply more of a commodity if the price is high (subject to capacity and
other constraints, of course). Conversely, the higher the price of a commodity, the
less likely people who use the commodity (the demand side) will want to buy it
and the more likely they will seek alternatives. This opposite response to the price
of a commodity means that there will be a price at which supply and demand are
perfectly balanced – the so-called market-clearing price. This is important because
supply of and demand for chemical commodities: (i) follows these same principles;
and, (ii) will fluctuate throughout time dependent upon many variables, including:
● raw material availability
● labor availability
● success or problems with production operations
● success or problems with operations, which consume the chemical
● logistics problems
● and many other factors
Figure 2.1 shows how this simple yet powerful economic theory works in an illus-
trative example.
Now that we have introduced the building blocks, the concept of commodity mar-
kets, and the principles of supply and demand, we can start to take a closer look at
the market situation with respect to our major building blocks of the modern petro-
chemical industry.
Supply
Price (example, dollars per unit)
Surplus
Market-claring price
at Q0, P0
Shortage
Demand
MTO/CTO – 4.0%
Gasoil – 2.7% Others – 9%
Global ethylene market – capacity and demand
Ethylbenzene – 5%
250 000 0.90
LPG – 15.3%
0.89 Ethylene dichloride – 9%
0.83
0.81
Naphtha – 40.4% 0.80
08
09
20
21
22
23
24
25
12
13
14
15
16
17
18
19
10
11
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
Capacity Demand Operating rate
Figure 2.2 Global supply and demand dynamics for ethylene. Source: Based on data from
IHS Markit and internal Honeywell UOP Estimates.
forward. Clearly, both the production of ethylene and the construction of ethylene
complexes are large and growing businesses.
As of 2019, the capacity to produce ethylene exceed the demand for ethylene by
about 12–13%. This is evident in Figure 2.2 by the difference in the size of the bars
representing capacity and demand. It is also shown in the line, which is the ratio of
demand to capacity and is frequently termed the operating rate. This is the global
operating rate needed for supply from the available capacity to meet the demand.
As such, it does not represent the operating rate at any individual plant; it is simply
a mathematical construct, which allows the balance of supply and demand. Despite
being grounded only in this theoretical balance, it is a useful number to watch.
Production plants of any kind, ethylene plants included, are rarely able to operate
at full rates for 365 days a year – plant upsets, weather, equipment breakdowns,
logistics challenges all conspire to reduce the theoretical limit to a practical limit.
As such, when the global operating rate approaches a critical value, it is frequently
a strong indication that more capacity – either through expanding existing facilities
or building new ones – must occur for supply and demand to remain balanced.
In the chemical industry, the “bellwether” operating rate value is typically around
90–92%. When operate rates are expected to exceed this value, supply disruptions
become more likely. As can be seen from Figure 2.2, the global operating rate for
ethylene has been rising for most of the past decade as the demand has recovered
from the global recession of 2008–2009, consuming some of the overcapacity, which
resulted from a construction boom in the Middle East [7] that occurred around the
same time. Our forward expectation is that the global operating rate will remain in a
tight band between 87% and 90%, which is a historically high level for the operating
rate. This is likely to mean regional supply pressures can occur more frequently due
to plant impairments as has been experienced in recent years [8].
The right-hand side of the Figure 2.2 illustrates the uses of ethylene. Production
of polyethylene in various grades dominates the demand for ethylene, with about
62% globally. A significant fraction of the polyethylene produced (over 50% by most
estimates) is used in nondurable customer packaging, including food packaging and
bottles; sadly, in the United States, less than 30% of this material was recycled in
2017 [9]. Single-use and nondurable uses of plastics were under significant scrutiny
in 2019 [10], and the continued trajectory for polyethylene is unclear. However, the
short-term offers few alternatives to the cost and convenience of polyethylene. It will
be interesting to see how the environmental and the economic challenges around
nondurable plastics are resolved in the coming years.
The production of ethylene for use in ethylene oxide manufacture follows at a
distant 15%. Ethylene oxide has three predominant uses: as a freezing-point depres-
sant and heat-transfer improver for mixtures in water in automotive antifreeze, in
surface-active nonionic surfactants, and as a key raw material in the production of
polyesters. The production of polyester represents the majority of the demand for
ethylene oxide. Ethylene dichloride (9%) and ethylbenzene (5%) account for most of
the remaining share of ethylene consumption, being used in the vinyls and styrenics
plastics chains, respectively. The remaining demand (9%) is spread across a number
of small products throughout the plastic and chemical value chains.
22 2 Petrochemical Markets
US Western Europe
Ethane Propane Butane Naphtha Gas oil M/CTO Other Ethane Propane Butane Naphtha Gas oil M/CTO Other
Ethane Propane Butane Naphtha Gas oil M/CTO Other Ethane Propane Butane Naphtha Gas oil M/CTO Other
Figure 2.3 Regional variation in feedstock sources for ethylene. Source: Figure created
using data adapted from IHS Markit [11].
Throughout this section, we have used global facts and figures to illustrate the fea-
tures of the market for ethylene. It is important to note, however, that the regional
markets for ethylene (and others of the building blocks) can be quite different from
geography to geography. While it is beyond the scope of this book to describe all of
these regional variations, we have constructed Figure 2.3 to illustrate the differences
that can occur. Figure 2.3 compares the sources of ethylene in four large produc-
ing regions – the United States, Western Europe, the Middle East, and Northeast
Asia – in 2016 [11]. The United States and the Middle East show large percentages of
ethylene produced from ethane because of the availability of the feedstock. Western
Europe and China show very low proportions from ethane, but quite high propor-
tions of ethylene produced from naphtha. In addition, China shows about 7% of its
ethylene produced from coal and methanol to olefins technology – a pathway not
used in any of the other regions. These significant regional differences lead to a rich
and varied global market.
At almost 114 million metric tons estimated to be produced in 2019 [12], propylene
is one of the top commodity chemicals produced as an intermediate for a large
variety of chemicals compounds. Because of the existence of both aliphatic and
olefinic bonds in the molecule, the chemical versatility of this molecule is sub-
stantial; indeed, propylene is the smallest such molecule with both key types of
carbon–carbon bonds. The chemical reaction’s versatility and ability to synthesize
so many unique compounds stem directly from the presence of these highly varied
bonds. Though not as concentrated as ethylene, more than 35% of the world’s
propylene production comes from the top 10 producing companies. Because of the
2.3 The Market for Propylene 23
High severity FCC/other – 4.0% Global propylene market – capacity and demand
MTO/CTO/MTP/CTP – 4.6% Acrylic acid/others – 9.5%
200 000 0.84
Metathesis – 3.8% Cumene – 5.9%
Propane dehydro (PDH) 12.1% 180 000 Oxo alcohols – 5.9%
0.82
160 000 Acrylonitrile – 5.6%
80 000
0.76
60 000
Polypropylene – 66.9%
40 000
0.74
Steam cracking – 46.8% 20 000
08 0.72
09
20
21
22
23
24
25
12
13
14
15
16
17
18
19
10
11
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
Capacity Demand Operating rate
2019 sources 2019 users
Figure 2.4 Global supply and demand dynamics for propylene. Source: Based on data from
IHS Markit and internal Honeywell UOP Estimates.
complexity and size of propylene production and markets, these companies also
tend to be the world’s largest integrated oil refining and petrochemical companies.
Propylene is marketed typically in three major grades – minor, more specialized
grades do exist – referred to as refinery, chemical, and polymer grade. They are listed
in order of increasing purity. Refinery grade contain typically between 50% and 75%
propylene, while chemical grade is typically 92–98% propylene. Polymer grade is
always equal to or greater than 99.5% propylene, but there are multiple different
impurity specifications (on methylacetylene and propadiene, for example), which
alter the grades slightly. Our standard view for the capacity and demand balance
for all propylene grades is shown in Figure 2.4. Starting from the left-hand bar, we
can see that the sources of propylene are much more varied than the sources of
ethylene. Only about 47% of the propylene produced in 2019 was through steam
cracking, as described in Chapter 1 and in Section 2.2. This compares to more than
95% of the ethylene produced coming from the same process. Indeed, propylene
is the primary “co-product” of the thermal conversion of hydrocarbons to make
ethylene. However, as hydrocarbon feedstocks for the production of ethylene shift
toward abundant ethane from heavier hydrocarbons, the amount of propylene
produced as a co-product declines. This is illustrated in Table 2.1, which presents the
relative production of key intermediates from the thermal conversion of different
hydrocarbon species.
A significant portion of the world’s sources of propylene comes from refineries.
The so-called fluid catalytic–cracking process, or FCC unit, which focuses on the
production of naphtha for motor fuels, also produces an off-gas byproduct, which
contains a substantial quantity of propylene. From Figure 2.4, we can see that
this source accounts for about 29% of the world’s supply of propylene. Adding
the production of propylene from FCC (byproduct of gasoline manufacture) and
steam cracking (byproduct of ethylene manufacture) shows that slightly more
than 75% of the world’s propylene results in recovery of this value material as a
byproduct of manufacturing routes primarily focused on other compounds (gasoline
and ethylene). This is one manner in which the markets become complex and
linked – for example, a producer who is focused on the ethylene market must also
24 2 Petrochemical Markets
Table 2.1 Relative amounts of byproducts, including propylene, from thermal cracking of
various feedstocks.
Products
Ethylene 1000 1000 1000 1000
Propylene 20 347 376 516
Butylenes — 64 126 320
Pyrolysis gasoline — 177 274 730
Fuel gas 150 719 623 813
watch and make assessments of the propylene market to understand the feasibility
of new capacity investment, or of increasing or decreasing operating rates.
The remainder of the source of propylene, slightly less than 25% of the world’s
supply, comes from processes, which are described as “on-purpose” propylene pro-
duction. As the name implies, these processes have the production of propylene as
their primary purpose. These processes have gained substantially in popularity in
recent years, driven by a separation in growth rate between propylene and ethylene.
The consumption of propylene has typically been growing at a slight multiple to the
production of ethylene – for example, in recent years, propylene demand growth
rates have been between 4.0% and 4.5%, while ethylene demand has been increasing
between 3.0% and 3.5%. Additionally, gasoline demand has been declining, resulting
in lower operating rates for FCC units and lower production of propylene. This has
put pressure on propylene markets as recently as 2012–2014, resulting in the need
for additional sources of propylene. These on-purpose methods are frequently seen
as more expensive routes, but often simplify the business of propylene as they result
in fewer (often times only one or two) products, which need to be marketed.
Most significant in the growth of these on-purpose methods has been propane
dehydrogenation (also called PDH) technology, which now accounts for more
than 10% of the world’s production of propylene. This process starts with propane,
an abundant and relatively cheap raw material, and makes only propylene as the
product, thereby simplifying the business operations and diversifying both the raw
materials and the production routes to propylene [13]. Significant new capacity
for PDH units has been built in China using important propane from the United
States and the Middle East. In addition to PDH, methanol to olefins (MTO) and
coal to olefins (CTO) units produce propylene besides ethylene. This production
technology has also seen a growth in adoption in China in recent years with its
large reserves of coal [14].
The middle portion of Figure 2.4 shows the global capacity and demand balance
(the bars) for propylene. The world’s capacity to produce propylene has increased
2.3 The Market for Propylene 25
steadily in recent years between 4.0% and 4.5%, with capacity reaching approxi-
mately 138.1 million metric tons in 2019 [12]. As with ethylene, capacity addition
is not perfectly smooth due to the size of the plants and the timing of construction
decisions; however, the capacity curve is much smoother owing to the smaller capac-
ity of typical propylene plants, particularly the on-purpose plants. As with ethylene,
the difference in the size of the bars for the capacity to produce and the demand for
propylene is accounted for by the operating rates of the production plants, shown in
Figure 2.4 as the line. The operating rate for propylene has remained quite constant
over the last 20+ years, from a low of about 76% during the global financial crisis
of 2008–2009 to a high of about 83–85% from 2004 to 2007. Because of the demand
growth and the relatively stable operating rate, the environment for new investment
in propylene production has remained strong and the market has remained quite
healthy over this same time.
Finally, the bar on the right-hand side of Figure 2.4 shows the approximate dis-
tribution of uses of propylene in 2019. Like with ethylene, the large majority of
propylene produced (66.9%) gets consumed for the production of various grades of
polymer, in this case, polypropylene. Other significant uses of propylene include the
production of propylene oxide (8%), acrylonitrile (6%), and a variety of other chem-
ical compounds, including cumene/phenol, oxo alcohols, acrylic acid, and superab-
sorbent polymers. While some polypropylene ends up in single-use plastics (bottle
caps, closures, some food packaging), a larger percentage of polypropylene ends
up in durable goods (carpets, other fibers, engineered plastics, etc.). This makes
polypropylene likely to be less subjected to the scrutiny placed on single-use plastic
materials than, for example, polyethylene.
The production of propylene has seen dramatic shifts in raw material uses and
geographic distribution in recent years, and these shifts are likely to continue [12].
Most of these shifts have been caused by observable and predictable macroeconomic
trends. As we have previously mentioned, the shift in feedstocks for the steam crack-
ing of hydrocarbons from heavier materials (e.g. naphtha) to lighter hydrocarbons
(e.g. ethane) has resulted in a lowering of the propylene produced as a byproduct
of ethylene (see Table 2.1). Adding to this decline in traditional supply has been a
decline in gasoline demand, which results in less propylene produced as a byproduct
from the FCC unit. Occurring simultaneously has been the broad industrialization
of China and the increased demand of the materials produced from propylene (e.g.
polypropylene) to be available in China. The resulting increase in production of
propylene in China has been enormous, as seen in Figure 2.5. Indeed, since 2013,
China has overtaken North America as the region with the highest capacity to pro-
duce and, by 2021, will be nearly double the production capacity of the next largest
region (see Table 2.2, [15]) – all occurring in a span of less than 10 years. Much of
this capacity expansion has been enabled by the combination of increased propane
supply from shale exploration in the United States as well as the recent boom in ship-
ping availability [16]. It is these types of shifts that make the petrochemical markets
dynamic, exciting, and, at times, frustrating.
26 2 Petrochemical Markets
2013 2019
North America
South America
Europe (WE+CE)
Russia and CIS
108.9 M MT/a 138.1 M MT/a
MEA
SE Asia + India
China
NE Asia example, China
Like the markets for ethylene and propylene, which are large and have widely var-
ied end products for which they are intermediates, the C4 olefins and dienes are
produced in slightly smaller quantities but also have large and widely varying uses.
However, there is one complication – while we group these C4 olefins and diolefins
for convenience, there are really three separate compounds we need to consider:
n-butenes, isobutene, and butadiene. While the end uses of the three separate com-
pounds are quite different, the sources of these compounds are often the same or at
least highly integrated as we will see in the discussion below. For simplicity, we will
split the three classes of compounds into two distinct markets: butylenes, comprising
normal and iso C4 olefins, and butadiene.
In 2019, the world’s capacity to produce butylenes was approximately 80.1 million
metric tons per annum, broken down into 46.3 million tons per year of n-butenes,
and 33.8 million tons per year of i-butenes [17]. These butylenes are produced
through three main refinery and chemical processes: FCC (as we saw with propy-
lene in Section 2.3), steam cracking (as with both ethylene and propylene), and the
direct conversion of isobutane to isobutene through the butane dehydrogenation
2.4 The Market for C4 Olefins and Diolefins 27
40 000
0.56
30 000
FCC, 45.4% 20 000 Alkylate, 43.6%
0.54
10 000
0.52
12
13
14
08
09
20
20 1
22
20 3
24
25
15
16
17
18
20 9
10
11
2
1
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
2019 sources 2019 uses
Capacity Demand Operating rate
process (analogous to the PDH process for propylene production). In 2019, the
capacity to produce butylenes was roughly estimated as 45.4% FCC, 18.4% steam
cracking, 15% BDH (butane dehydrogenation), and 21.2% from a variety of other
smaller chemical processes (as byproducts) [17]. This information on the capacity
to produce butylenes is summarized in Figure 2.6.
The demand for butylenes is quite a bit smaller than the world’s capacity to
produce, with 2019 seeing a global butylene demand of approximately 49.0 million
metric tons [17]. This demand is split nearly equally between normal (23.9 million)
and isobutenes (25.1 million). Interestingly, we can also segment the butylene
demand into demand for fuels (approximately 39.7 million metric tons in 2019)
and demand for further petrochemical production (approximately 9.4 million
metric tons). Butylenes are used in fuels in predominantly two different ways. The
olefins can undergo alkylation or dimerization chemistry to produce a hydrocarbon
product of longer chain length and a higher degree of branching. This results in
a higher-octane value for the molecule, which is beneficial for gasoline internal
combustion engines [18]. The second method involves reacting the olefins with an
alcohol to make an ether, for example, methyl tertiary butyl ether (MTBE), ethyl
tertiary butyl ether (ETBE), or tertiary amyl methyl ether (TAME). These ether fuel
additives are prized for both their octane values, which are typically higher than
pure hydrocarbons, but also because they add oxygen to the gasoline, which has
been shown to improve combustion leading to lower levels of air pollutants [19].
In addition to fuel uses, the butylenes are used as a raw material for the pro-
duction of a number of chemical compounds, including propylene (through mixed
ethylene/butylene metathesis), 1-butene (for polymer additives), monomethyl acry-
late (MMA), polyisobutylene (PIB), sec-butyl alcohol/methyl ethyl ketone (MEK),
to name a few. Because the end use typically has a specific chemical requirement
(for example, MTBE is made from isobutylene, not n-butylenes), the estimation of
a capacity utilization is quite complex and depends strongly on the process and the
ratio of n- and isobutenes produced by the process. As a result, we have shown both
the fuel and chemical demand levels and the n- and isobutene demand breakdowns
on the right-hand side of Figure 2.6 [17].
It is tempting to attempt to separate out the different butylene compounds and do
a market balance on each butylene component. While this can be approximated at
28 2 Petrochemical Markets
Table 2.3 Butylenes produced from steam cracking and refinery sources (thousands of
metric tons per annum) for selected years.
the aggregate level, it is difficult to break down the specific sources of butylenes. For
example, what is the number of metric tons per annum of n-butene produced from
steam cracking? The answer to this, even at an approximate level, would require a
very detailed analysis involving factors (at a minimum) such as the feedstock and
capacity of all the steam-cracking units, and the extent to which the units recover
the butylenes produced. This is well beyond the intention of this chapter. Instead,
we can look at the aggregate production of butene isomers from the various sources
and derive some conclusions.
Table 2.3 shows the reported annual production of the two primary categories
of butylenes, n-butene and isobutene, for selected years (2000, 2005, 2010, 2015,
and an estimate for 2020) from both steam-cracking operations and from refiner-
ies [20]. A couple of interesting trends can be observed. First, the amounts of C4
olefins declined from 2000 to 2010 and have since begun to rebound. This is true
for both FCC and steam-cracking sources. The decline in C4 olefin production in
steam cracking corresponds to the period of time of a large shift in feedstock eco-
nomics favoring lighter feeds like ethane and propane. During this period, despite
increasing supply of ethylene from steam cracking, the amount of C4 olefins pro-
duced declined because the feedstocks got much lighter on average. However, once
all the units, which can change to lighter feedstocks, have done so, then additions
of ethylene production also mean additions of C4 olefin production, and this is seen
in the later years. For the FCC units, the production numbers were quite stable dur-
ing the period of 2000–2010 and then began to increase substantially from 2010 to
2020. The increase since 2010 corresponds to a period of time when alkylate for fuels
became increasingly valuable from refineries due to global clean fuel standards [21]
resulting in more alkylate (and therefore C4 olefin) production.
Our second observation is that the ratio of normal to iso olefins resulting from
the steam-cracking operation is much more consistent over the years shown than
the ratio from the FCC unit. Steam cracking is a thermal process and there are
limited choices to be able to adjust the ratio of normal to iso, and largest among this
is the feedstock. As a result, the ratio remains much more stable as not all plants
can adjust feedstock readily. As discussed above, the shift to lighter feedstocks is
2.4 The Market for C4 Olefins and Diolefins 29
consistent with the slight increase in normal to iso ratio toward the end of the
period shown. The ratio of normal to iso olefin produced from the FCC process
from 2000 to 2020 varies widely from a low of 0.937 to a high of 1.565. There are
likely multiple reasons for this observation, but one involves changes to the FCC
process. The FCC process uses a catalyst, one which has been engineered over
the last several decades to produce more isobutane and isobutene as the branched
isomers are more valuable to the refinery for alkylation. This is most likely the
reason for the significant shift from 2000 to 2010.
It is important to note that there is currently substantial overcapacity for the
production of butylenes, but it is not homogeneously distributed between n-butenes
and isobutene. The demand for n-butenes corresponds to a utilization of capacity of
slightly higher than 50%, while the same estimate for isobutene shows a utilization
of close to 75%. Despite isobutene being the less thermodynamically stable isomer,
there is a stronger demand growth for isobutene due to its significant use across
many compounds – alkylate, fuel additives, and chemical uses – than there is for
n-butenes.
Although produced in much lower volume than butylenes, butadiene is a much
more valuable compound used exclusively in chemical production. In 2019, it is esti-
mated that global demand for butadiene will reach almost 13 million metric tons
against a capacity of approximately 13.4 million metric tons. The production of buta-
diene is quite tight – meaning that producing assets operate a quite high utilization
level. Indeed, the utilization of butadiene production assets runs typically in the
range of 90–95% (see Figure 2.7), which is quite high for petrochemical produc-
tion assets. There are several reasons for this. First, butadiene is produced almost
exclusively as a byproduct of the production of ethylene in crackers; butadiene is
only economically recovered from heavy feedstocks used in crackers (e.g. naphtha
and heavier feeds). As a result of being a byproduct and being a sought-after mate-
rial, the production of butadiene is quite lucrative, resulting in asset owners want-
ing to operate the assets as much as possible. Second, the crackers, which provide
the source, also have typically a very high utilization ratio because these are very
capital-intensive operations and return on investment requires high utilization.
While the only appreciable source of butadiene currently available is as a byprod-
uct from naphtha and heavier feed crackers, there are multiple uses of butadiene
and multiple competitive options, depending on the end market. Butadiene demand
growth has been rather unsteady over the past decade with historical year-over-year
growth varying from 1% to almost 5% from 2008 to 2018. Key drivers for this are
the aforementioned tightness of the supply. Consumers view this tightness as risk
and therefore look for alternative raw materials, which can be substituted, and this
substitution creates a rather “lumpy” demand profile. Tight supply also increases
the price at which people sell butadiene in the market, which opens up opportuni-
ties for competition as well. In addition, one of the key substitute products (natu-
ral rubber, see below) is an agricultural product promoted by several governments
in Southeast Asia in recent years as a means for rural economy improvement. In
Figure 2.7, the forward growth curve for butadiene has been estimated at about 4%
30 2 Petrochemical Markets
14.00
ABS, 12%
10.00
SB rubber, 27%
8.00 0.85
6.00
0.75
2019 uses
12
13
08
09
14
20
21
22
23
24
25
15
16
17
18
19
10
11
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
Capacity Demand Operating rate
growth year-over-year, but this will likely vary substantially for the same economic
and competitive reasons mentioned above.
Butadiene is used as raw material for a number of engineered polymers and elas-
tomers. Figure 2.7 also shows the global demand by use of butadiene. Note that
there have been substantial shifts over relatively short periods of time in the uses.
Butadiene rubber and styrene-butadiene rubber (SBR) remain the largest uses, sup-
porting 30% and 27% of the global demand, respectively, in 2019. However, over the
last 10 years, the consumption toward butadiene rubber has overtaken SBR because
of increased uses of butadiene rubber in high-performance tires for the automotive
market [22]. Within tires, butadiene rubber is always used in conjunction with SBR
and often with natural rubber, an agricultural product from the rubber tree [23].
Indeed, adjustments to the produces’ formulae for production of tires often change
as the relative cost of synthetic polymers and natural rubber changes. Other sig-
nificant uses for butadiene rubber include footwear, wire and cable insulation, and
conveyor belts [22]. Besides SBR and butadiene rubber, butadiene is key raw mate-
rial for the production of a co-polymer known as acrylonitrile-butadiene-styrene, or
ABS. This is rapidly growing engineered polymer used for high-performance appli-
cations such as household appliances, automotive parts and interiors, and other
demanding applications.
The recovery of butylenes and butadiene from refineries and steam crackers is an
elegant, complicated, and highly integrated process. Figure 2.8 shows a simplified
generic flow scheme for the recovery processes resulting in butadiene, isobutylene,
and n-butylenes and a few critical products of the C4’s as well. Crude unsaturated C4
streams from steam crackers and refinery sources are shown on the left along with
isobutane, which can be converted into isobutylene through a dehydrogenation
process similar to that described early for on-purpose propylene production. In
general, only the unsaturated C4 streams from steam crackers will have a high
enough concentration of butadiene to warrant recovery, as shown in Figure 2.8.
When butadiene recovery is warranted, it is passed to a butadiene extraction unit,
2.4 The Market for C4 Olefins and Diolefins 31
Isobutane
Butane dehydro
High-purity isobutylene
FCC-based
crude C4ʹs Raff-1
Selective
hydrogenation Etherification
Steam cracker-
based crude C4ʹs
Acetylene Polyisobutylene
hydrogenation
which uses solvent extraction technology to selectively absorb the butadiene, and
then strip the butadiene from the solvent using heat [24]. Licensors of this process
technology include BASF, McDermott, Nippon-Zeon, LyondellBasell, and others
[25]. If butadiene is not desired, or not present in high enough concentration to
be economically recovered, the unsaturated C4 stream is typically passed through
a selective hydrogenation unit, which will convert the diene and acetylenes into
mono-olefins. The effluent from the selective hydrogenation units will now look
very much like the effluent from the butadiene extraction units. The industry has
termed this stream “Raffinate-1” or “Raff-1.” This Raff-1 stream will be a combi-
nation of isobutylene, normal butylenes, and butanes (saturated C4 compounds).
It is quite different from the effluent of an isobutane dehydrogenation unit, which
is predominantly isobutylene and butanes, but for the purposes of Figure 2.8, the
streams can be considered together from this point forward.
The Raff-1 stream then will generally pass to an isobutylene utilization or recovery
section. Isobutylene can be used to make MTBE (by condensation with methanol),
which is a high-octane gasoline additive used in many countries, or to make PIB.
If ultra-high purity isobutylene is desired, one can make MTBE first to recover the
isobutylene from the Raff-1 mixture, and then crack the MTBE back into isobuty-
lene and methanol. This reaction is equilibrium limited, but by separation of the
isobutylene you can recycle the MTBE to extinction and have high recovery of very
high isobutylene. This is frequently practiced to derive high-quality butyl rubber.
Following the recovery or consumption of the isobutylene, the industry has termed
this effluent stream as “Raffinate-2” or “Raff-2.” This stream is now predominantly
normal butylenes contained in saturated butanes and can be used further processed
to make pure normal butylenes, alkylate for motor fuel applications, solvents, or
other products [17]. There are a number of different options and processes, which
can utilize these unsaturated C4 streams, and Figure 2.8 should be considered
only one example configuration. The versatility of these unsaturated C4 streams
results in a wide number of important specialty compounds with critical uses in our
economy.
32 2 Petrochemical Markets
100.000 0.840
0.830
Ethylbenzene, 50.1%
50.000 0.820
Reforming, 39.1%
0.810
0.800
08
09
20
21
22
23
24
25
12
13
14
15
16
17
18
19
10
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
Capacity Demand Operating rate
transalkylation process. Extraction of benzene from liquids resulting from the cok-
ing process of coal used for steel manufacturing remains a significant contributor to
the capacity to produce benzene (8%). The remainder of benzene production capac-
ity is attributed to a process called hydrodealkylation (HDA), accounting for about
5% of global capacity. In the HDA process, toluene or (rarely) other alkyl-substituted
aromatic compounds are converted – usually in a thermal process – into methane
and benzene. This capacity usually runs intermittently as the economics of the con-
version of toluene into benzene are frequently not favorable.
The fraction of global capacity production given in the paragraph above represents
the total capacity amounting to a possible production of approximately 65.55 million
metric tons of benzene in 2019. In reality, only about 49 million metric tons of ben-
zene were produced. Figure 2.9 shows the sources of supply of benzene in 2019, and
the inset table in Figure 2.9 shows the approximate operating rate for the five dif-
ferent key sources of supply (e.g. the actual production divided by the capacity to
produce) in 2019. As mentioned above, the lowest operating rate is observed for the
swing capacity, HDA-based benzene, at about 31%. The highest estimated operating
rate in 2019 is observed for the production of benzene as a byproduct of p-xylene
production through the transalkylation process (almost 89%). This is because, as a
byproduct, production economics are less dependent on benzene alone. Indeed, the
amount of benzene produced by a p-xylene production facility has decreased dramat-
ically over the years driven by economics strongly favoring p-xylene over benzene.
The total production of benzene in 2019 was approximately 48.97 million met-
ric tons. The production of benzene has been increasing at an average rate close
to 2% for nearly a decade and this growth rate is expected to continue to the next
several years [28]. Benzene is used for a wide variety of end products, including
a number of monomers, which are essential for production of polymers such as
styrenics, phenolics, and nylon. Almost 50% of the world’s demand for benzene is
driven by the need to produce ethylbenzene (through benzene alkylation with ethy-
lene [28]); ethylbenzene is a critical precursor for the production of styrene, which,
in turn, is dehydrogenated into styrene – the starting monomer for polystyrene and
related styrenic materials. The next largest demand for benzene is through the pro-
duction of cumene (about 21%), which is similarly produced by the alkylation of
34 2 Petrochemical Markets
benzene with propylene [28]. Cumene is further converted into phenol and acetone
through an oxidation process [30]. Phenol is a critical monomer for the production
of the phenolic polymer chain, which includes phenolic resins (through conden-
sation with formaldehyde), nylon through hydrogenation to cyclohexanone, poly-
carbonates, and bisphenol-A through condensation with acetone. All these plastics
compounds are in wide use in household products as well as engineered materials.
Cyclohexane is a key intermediate in the production of adipic acid and caprolactam,
both of which are critical starting materials for the production of various grades of
nylon fibers [28]. In 2019, it is estimated that about 11% of the world’s benzene will
be hydrogenated to produce cyclohexane for these nylon polymers. Together, these
three value chains (styrenics, phenolics, and nylons) represent about 82% of the
demand for benzene production. The remainder of the benzene produced is used
in the synthesis of nitrobenzene (10%, intermediate in polyurethane), chloroben-
zene (2%, dyes and herbicides), maleic anhydride (2%, coatings and surfactants), and
other key intermediates for the vast petrochemical industry. The market for benzene
is described in Figure 2.9 using our common format.
0.620
21
08
09
10
11
12
13
14
15
16
17
18
19
20
22
23
24
25
20
20
20
20
20
20
20
20
20
20
2019 sources 2019 users
20
20
20
20
20
20
20
20
Capacity Demand Operating rate
of ethylene being the second largest source at about 18% [31]. The remainder of the
produced toluene comes from minor processes like extraction from coking ovens
and coal-based chemical processes.
The demand for toluene was approximately 27.8 million metric tons in 2019, grow-
ing at a rate between 2% and 3% annually [31]. This growth rate is expected to con-
tinue for the next several years as p-xylene demand continues to increase at about
4.5–5.5% growth per year. While toluene and the transalkylation reaction are a sig-
nificant source of xylenes for the production of p-xylene (see Chapter 5), most of the
new complexes being developed globally are moving toward net-zero isolation of
toluene [32], resulting in a slower growth rate for toluene demand. Also, large-scale
blending of toluene for gasoline property (vapor pressure and octane) improvement
is already practiced, and this has two significant consequences. First, the demand
growth for gasoline blending will be muted as most regions are now using toluene,
leaving growth equivalent of the overall gasoline demand growth only. Second, this
growth in gasoline blending has set a floor or minimum price for toluene in the mar-
ket as producers can simply sell toluene for gasoline blending. This has closed off a
traditional market of p-xylene production from toluene and mixed xylenes. Com-
plexes used to be able to make acceptable margins by purchasing toluene and mixed
xylenes on the market and converting them into p-xylene. In the past 10 years, this
has become a very difficult and low-margin business.
Toluene is isolated and produced for three basic reasons: the production of
benzene and xylenes (BX; usually related to p-xylene production); the production
of chemicals from toluene; and, gasoline blending. The largest demand of these
three is for the production of benzene and xylenes (BX), with demand reaching
about 13.1 million metric tons in 2019 [31]. The next largest demand is for blending
into gasoline, which reached approximately 7.8 million metric tons in 2019 [31],
though demand growth will slow in the coming years as indicated above. The final
category of toluene demand, chemicals production from toluene, includes a diverse
set of compounds, including toluene diisocyanate, benzoic acid, and common and
specialty solvents. This category accounted for about 7.0 million metric tons of
toluene demand in 2019 [31]. Over the next decade, it is likely that the highest
demand growth will remain aligned with the demand for p-xylene, while growth of
toluene-derived chemicals will start to approach a similar growth rate.
36 2 Petrochemical Markets
Other, 0.4% Global xylene market – capacity and demand Isophthalic acid, 1.0%
Pygas, 4.5% 120.00 0.800 Others, 1.4%
Solvents, 9.6%
0.780
100.00 Phthalic anhydride, 5.1%
Transalkylation, 25.6% 0.760
0.720
60.00
0.700 PTA/PET,polyester, 82.9%
Reforming, 69.5% 40.00 0.680
0.660
20.00
0.640
0.620
21
22
08
09
10
11
12
13
14
15
16
17
20 8
19
20
23
24
25
1
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
2019 sources 2019 uses
Capacity Demand Operating rate
Mixed xylenes come from a variety of sources, the most abundant of which
are naphtha reforming, pyrolysis gasoline from naphtha crackers, transalkylation
processes, and other sources, including coal-derived chemicals [33]. Almost 70%
of the world’s supply of mixed xylenes comes from naphtha reforming, with much
of this being on-purpose or petrochemical-oriented naphtha reforming. Some
continue to be extracted from motor fuel reformers, though this is becoming less
and less attractive due to the demands on octane in motor fuels. Only about 4–5%
of the mixed xylenes are supplied from naphtha crackers, and this percentage is
likely to decline over the next decade [33]. The increased demand for xylenes, and
the shift to lighter (ethane) feedstocks in crackers, virtually assures the importance
of pyrolysis gasoline will decline over the next decade. It will, however, remain
very relevant from specific geographic regions. About 25–30% of the xylenes are
supplied from the various transalkylation processes, which involve moving alkyl
groups, specifically methyl groups, across the aromatic rings. This will remain
roughly steady, and will grow with the naphtha-reforming capacity; in a modern
xylene complex today, it is very rare to build reformers without a transalkylation
unit.
As mentioned in an earlier paragraph of this section, the great majority of mixed
xylenes produced (83%) are used to make p-xylene, which, in turn, is used to make
purified terephthalic acid and, ultimately, polyester products (fibers, films, and bot-
tles). Slightly more than 5% of the demand for mixed xylenes is for the production
of o-xylene, a key raw material in the manufacture of phthalic anhydride (PA). PA is
uses in many PVC (polyvinyl chloride)-based products as a plasticizer. About 1% of
the demand for mixed xylenes is driven by m-xylene, despite the fact that this is the
isomer available in highest quantity when the xylenes are at equilibrium. m-Xylene
is used to add ductility and impact resistance to plastic bottles [33]. All of this means
the demand for mixed xylenes has been growing 3.5–4.5% per year and will con-
tinue to do so as long as polyester demand growth continues. In 2019, the demand
for mixed xylenes was approximately 60.2 million metric tons per annum [33], with
about 60% of this mixed xylene being consumed in Northeast Asia.
Synthesis gas is a widely used fuel-gas mixture, which is primarily composed of car-
bon monoxide and hydrogen [35, 36] and has primary uses in power generation,
synthesis of liquid fuels, use as a gaseous fuel, separated and used as industrial gases,
and for the synthesis of chemicals. Our primary focus in this section is on the use of
synthesis gas for chemical synthesis, which is slightly larger than 50% of the over-
all synthesis gas market in 2019. While there are many different chemicals that are
derived from synthesis gas, there are two chemicals that make up the vast major-
ity of the synthesis gas chemicals industry – methanol and ammonia. We focus our
discussion on these two chemicals in Sections 2.6.1 and 2.6.2.
38 2 Petrochemical Markets
0.780
Other uses (mixed), 13.8%
200.00 0.760
150.00 0.740
Natural gas reforming, 70.6%
0.720
100.00 Urea, 52.5%
0.700
50.00
0.680
0.660
20 1
20 2
09
10
08
11
12
13
14
15
16
17
18
19
20
23
24
25
2
2
20
20
20
20
20
20
20
20
20
20
20
20
20
20
America. Most of China’s ammonia capacity is based on coal as the hydrogen source,
and this capacity is coming under more and more environmental scrutiny.
The demand for ammonia has increased steadily across the last two decades at
about 2.0–2.5% with demand reaching 190.9 million metric tons per annum in 2019
[37, 38]. About one-third of the demand is driven by China, consistent with its
capacity – ammonia and ammonia-derived chemicals is one area where China has
achieved a strong level of self-sufficiency. At least 81% of the ammonia demand is
driven by fertilizers, which should be connected to population growth and increased
nutritional needs [37]. Slightly more than 50% of the ammonia produced ends up as
urea products for use directly or indirectly as fertilizers [36, 37]. After urea, there are
a number of smaller ammonia-based products like phosphate, nitrate, sulfate, etc.,
which share fertilizer and other markets as end-use applications. There are dozens
of other applications for ammonia, from refrigerants to cleaners to fermentation
feed, but the consumption for these uses is minor [39].
80.00 0.640
Formaldehyde, 26.8%
60.00 0.620
Natural gas reforming, 60.2%
40.00 0.600
Fuel uses, 28.8%
20.00 0.580
0.560
20 1
20 2
09
10
08
11
12
13
14
15
16
17
18
20 9
20
23
24
25
2
2
1
20
20
20
20
20
20
20
20
20
20
20
20
20
20
20
were formaldehyde and a fuel oxygenate additive called MTBE at about 7.7 million
metric tons and 5.9 million metric tons, respectively. Formaldehyde is used to
produce resin products for the wood products and building products industry [40].
Since the late 1990s, despite a ban on MTBE use in the United States, the fuel usage
of methanol has grown dramatically reaching 25.7 million metric tons in 2019 [42].
This comes in the form or MTBE, another ether called tertiary amyl methyl ether
(TAME), dimethyl ether, and direct fuel blending; in 2019, direct fuel blending
into gasoline had reached almost 7 million metric tons, from 0 in the late 1990s
[41]. In addition, a new market for methanol – converting methanol into ethylene
and propylene – has emerged in China, consuming over 13 million metric tons
of methanol in 2019 [41]. In addition, formaldehyde use has continued to grow,
reaching 22.6 million metric tons per annum in 2019. Considering all these factors,
and the ups and downs of various different end uses, the demand for methanol is
estimated to be just slightly above 84 million metric tons in 2019 [41]. Since 2008,
the demand has nearly doubled in 11 years for a 6.4% annual growth rate.
Looking at the split of demand by use in 2019, we find that about 27% of the
demand is driven by formaldehyde and building products end uses, which is the
largest single chemical use. However, more than 28% of the methanol produced ends
up in the fuels sector through MTBE and other fuel ethers, direct gasoline blending,
and the biodiesel industry. The fastest growing use, and third largest, is the MTO
application (about 19%), where methanol is converted to ethylene and propylene
(see Chapter 3). The remainder of the methanol is used in acetic acid (8%), solvents
(4.1%), and many other small chemical compounds [42]. This is summarized in the
capacity-demand figure for methanol in Figure 2.13.
References
9 Toto, D. (17 December 2018). Plastic bottle recycling rate declines in 2017.
https://www.recyclingtoday.com/article/2017-plastic-bottle-recycling-rate/
(accessed 04 September 2020).
10 Loepp, D. (30 May 2019). Under pressure, plastics find place in the global circu-
lar economy. https://www.plasticsnews.com/blog/under-pressure-plastics-find-
place-global-circular-economy (accessed 04 September 2020).
11 CEH Ethylene. IHS Markit.
12 (2017). CEH Propylene. IHS Markit and internal Honeywell UOP estimates.
13 Myers, R.A. (2018). Handbook of Petrochemical Product Processes, 2e.
McGraw-Hill, Section 10.3.
14 EChemI Website (12 July 2018). 2018 China newly added methanol and
MTO/CTO capacity statistics. http://www.echemi.com/cms/13515.html (captured
08 November 2019).
15 (2017). CEH Propylene. IHS Markit.
16 Rekoske, J.E. (2016). Technoeconomic impacts of abundant natural gas liq-
uids on the chemical industry. Presentation at the 11th Natural Gas Conversion
Symposium, Tromso, Norway (6 June 2016). Natural Gas Conversion Board:
Washington, DC.
17 (2017). CEH Butylenes. IHS Markit and internal Honeywell UOP estimates.
18 U.S. Department of Energy, Energy Information Agency (13 February 2013).
Alkylation is an important source of octane in gasoline. https://www.eia.gov/
todayinenergy/detail.php?id=9971 (captured 08 November 2019).
19 U.S. Department of Energy, Energy Information Agency Document. MTBE, oxy-
genates and motor gasoline. https://www.eia.gov/outlooks/steo/special/pdf/mtbe
.pdf (captured 08 November 2019).
20 CEH Butylenes. IHS Markit.
21 Kranz, K., duPont STRATCO Alylation Training. (September 2008). Intro to
alkylation chemistry. https://silo.tips/download/intro-to-alkylation-chemistry.
22 Encyclopedia Britannica. Butadiene rubber. Online Edition. https://www
.britannica.com/technology/butadiene-rubber (captured on 23 October 2019).
23 Sethurai, M.R. and Mathew, N. (eds.) (1992). Developments in Crop Science:
Volume 23, Natural Rubber. Elsevier Science.
24 White, W.C. (2007). Chemico-Biological Interactions 166 (1–3): 10–14.
25 Nizamoff, A.J. (2013). On-Purpose Butadiene. Nexant ChemSystems PERP
Report.
26 Carey, F.A. and Sundberg, R.J. (1990). Advanced Organic Chemistry Part A:
Structure and Mechanisms, 499. New York: Plenum Press.
27 Spitz, P.H. (1988). Petrochemicals: The Rise of an Industry, 42. New York: Wiley.
28 (2017). Benzene World Analysis. IHS Markit.
29 (2018). Mixed Xylene. World Economics Handbook, 15. IHS Markit.
30 Matar, S. and Hatch, L.F. (1994). Chemistry of Petrochemical Processes, 265. Lon-
don: Gulf Publishing.
31 (2018). CEH Toluene. IHS Markit and internal Honeywell UOP estimates.
32 Internal Honeywell UOP analysis.
42 2 Petrochemical Markets
33 (2018). Mixed Xylene. World Economics Handbook. IHS Markit, and internal
Honeywell UOP estimates.
34 Myers, R.A. (2018). Handbook of Petrochemical Product Processes, 2e.
McGraw-Hill, Section 13.3.
35 Chauvel, A. and Lefebvre, G. (1989). Hydrogen, synthesis gases and their
derivatives. Petrochmical Processes, Part 1: Synthesis Gas Derivatives and Major
Hyxrocarbons, 27. Paris: Editions Technip.
36 Speight, J.G. (2019). Natural Gas, A Basic Handbook, 2e. London: Gulf Profes-
sional.
37 (2017). CEH Ammonia. IHS Markit.
38 Internal Honeywell UOP estimates.
39 Appl, M. (1999). Ammonia: Principle and Industrial Practice. Wiley-VCH Verlag
GmbH.
40 (2017). CEH Methanol. IHS Markit.
41 (2017). CEH Methanol. IHS Markit and internal Honeywell UOP estimates.
42 Olah, G.A., Goeppert, A., and Prakash, G.K.S. (2018). Beyond Oil and Gas: The
Methanol Economy, 3e. Wiley-VCH.
43
Section II
3.1 Introduction
Olefins are key building blocks of chemical industry. Light olefins, ethylene and
propylene, are the most important olefins used in the manufacture of a large number
of chemicals in commodities and specialties such as polyethylene, ethylene dichlo-
ride, ethylene oxide, polypropylene, propylene oxide, and acrylonitrile. Butylene
is also in high demand for further processing to produce methyl-tert-butyl ether
(MTBE) and alkylate for gasoline blending. Heavy normal-olefins (C10 –C14 ), derived
from straight run kerosene, are the main feedstock component of linear alkylben-
zene (LAB), the most common raw material in the manufacture of biodegradable
household detergents. In this chapter, the manufacturing technologies having com-
mercial significance for producing these olefins will be discussed, including steam
cracking, paraffin (light paraffin and heavy normal-paraffin) dehydrogenation, and
methanol to olefins. There are other processes in a refinery, e.g. fluid catalytic crack-
ing (FCC), that produce light olefins as byproducts. These technologies will not be
discussed in this chapter.
Synthetic gas (syngas), a mixture of carbon monoxide and hydrogen, is used in
large volumes for the production a wide variety of commodity chemicals such as
hydrogen, methanol, ammonia, olefins, and MTBE. Hence, it represents a suitable
alternative to produce base chemicals from resources other than crude oil. Syngas
manufacturing technologies discussed in this chapter include conversion of natural
gas to syngas and coal gasification to syngas.
Ethane Ethylene
Gas propane
butanes Propylene
Gaseous
Naphtha Crude C4’s: products
Liquid
gas oils (Butadiene and
butylene)
Steam cracking
furnace Pyrolysis
800 – 900°C gasoline:
Low pressure (Aromatics
and C5’s) Liquids
No catalyst
Heavies
There are three steps in the steam-cracking process: pyrolysis and cooling, com-
pression and acid gas removal, and cryogenic cooling and product separation.
Pyrolysis and cooling: The feedstock is initially mixed with a dilution steam and
heated to 815–900 ∘ C at 1.7–2.4 bars. The resulting products are immediately
cooled to 340–510 ∘ C and further cooled to 37–43 ∘ C in a water quenching tower.
Compression and acid gas removal: The cracked gas stream is compressed from atmo-
spheric pressure to 35–38 bars. During the compression, acid gases (H2 S and CO2 )
are removed by scrubbing the stream. The resulting stream is dried over a solid
adsorbent.
Cryogenic cooling and product separation: The dried gas is cooled to between −90
and −130 ∘ C in a refrigeration train. The resultant gas, consisting primarily of
hydrogen, methane, and carbon monoxide, is purified to recover hydrogen; the
remainder is burned as plant fuel. Low-temperature fractionation separates the
dried products. Ethane and propane are generally recycled, while hydrogenation
of acetylene, propyne, and propadiene yields ethylene and propylene.
For the conventional endothermic cracking process, ethylene yield is improved by
raising the cracking temperature and reducing residence time. The operating condi-
tions are constrained by the metallurgy of the furnace tubes and coking tendency in
the cracking coils. The formation of coke deposits in cracking furnace coils remains
one of major obstacles in steam-cracking operations. Coke formation and subse-
quent accumulation in furnace components result in significant energy and main-
tenance costs. Typically, a naphtha pyrolysis furnace is decoked every 15–40 days.
Maximum cycle time is 60–100 days.
Ethylene fractionation separates ethylene as a high-purity product (>99 wt%) from
ethane, which is combined with propane and recycled for cracking. Propylene is
separated from propane in the propylene fractionation section as a chemical-grade
product (typically 93–95 wt% min) or more frequently as a polymer-grade propylene
(>99.5 wt%).
A mixed C4 stream is produced from the cracking of naphtha and heavy feedstocks
providing an important source for butadiene and butylene. More than 95% of world
3.2 Steam Cracking 47
Table 3.1 Typical products from cracking various feedstocks in a 450 KMTA ethylene plant
(KMTA).
Vac
Products Ethane Propane N-Butane Naphtha gas oil
supply of finished butadiene is extracted from the mixed C4 produced in the steam
crackers. The production of mixed C4 stream is strongly dependent on cracking
severity and feedstock specification. Typically the amount of mixed C4 produced
per ton of ethylene will decrease when running the cracker at higher severity.
The selection of feedstock for steam cracking is considered as one of the most
important issues for a petrochemical complex. Steam crackers are designed for feed
flexibility so that they can process a predefined range of feedstocks. Product depen-
dence on feedstock is shown in Table 3.1.
An ethane cracker is preferred when the feedstock is available in sufficient
amounts. As a result of shale gas development in the United States producing high
abundance of ethane, ethane cracking has become more attractive due to its low
cost, but byproduct credits from naphtha-based crackers in Europe and Asia could
be enough to keep them profitable maintaining a large share of the olefin market [1].
In a naphtha-based cracker, the feedstock contains normal-paraffins, isoparaffins,
naphthenes, and aromatics. Ethylene and other olefins are formed primarily from
paraffins and naphthenes in the feedstock. Normal-paraffins are the preferred com-
ponents for high ethylene and propylene yields. Isoparaffins produce much smaller
yields of ethylene and propylene than normal-paraffins.
An option to maximize the profitability from a naphtha-based cracker is to inte-
grate it with a refiner by using UOP MaxEneTM . The MaxEne process can simul-
taneously increase the yield of high value products from a naphtha-based cracker
and catalytic reforming unit. This process is designed to extract normal-paraffins
from full-range naphtha, consisting of molecules in the C6 to C10 range. The separa-
tion uses a simulated countercurrent contacting scheme, in which normal-paraffins
are first adsorbed in a series of fixed beds of adsorbent and then desorbed using a
desorbent. The extract stream with high-purity normal-paraffins makes a superior
48 3 Olefins and Synthesis Gas Production Technologies
n-Paraffins
Steam 30% Higher
Naphtha cracking ethylene yield
( C6–C10)
MaxEne
steam cracker feed, while the raffinate stream enriched with naphthene/aromatic is
optimal as a reforming feed (Figure 3.2). The process enables:
● 30% ethylene yield increase from existing naphtha cracker with no loss in propy-
lene,
● 5–6 RON increase of reformate from reforming unit to gasoline pool, and
● 2–3% aromatics yield increase for reformate from reforming unit to aromatics com-
plex.
In the late 1980s, the application of chromium catalysts with cyclical operation was
extended by Houdry to the dehydrogenation of propane to propylene and isobutane
to isobutylene. The new process application was called CatofinTM [2].
In 1959, a fluidized-bed dehydrogenation technology with chromium catalyst
was developed and commercialized in the former Soviet Union. This reactor design
avoids the cyclical operation by circulating the catalyst from the reactor to the
regenerator. Circulating regenerated catalyst is used to provide the heat of reaction
in the riser, while spent catalyst is reheated by carbon burn in the regenerator.
The technology was commercialized by Yarsintez [3] and further developed and
commercialized by Snamprogetti [4, 5].
The Philips STAR process also features regeneration of the catalyst on a cyclic basis
using a fixed-bed fired-tube reactor. Steam is used to lower the partial pressure of
the hydrocarbons and to slow down the deposition of coke on the catalyst providing
for extended cycle times. For continuous operation, various furnace modules (banks
of tubes) can be operated in parallel with several modules operating in the process
mode, while one is in the regeneration mode.
In the early 1970s, UOP introduced continuous catalyst regeneration (CCRTM )
technology that enabled noble metal catalysts to remain at their most desirable
activity for several years without having to shut down the reactor for catalyst
regeneration. The combination of noble metal catalysts operating at high severity in
conjunction with CCR technology made it possible to design, build, and economi-
cally operate large catalytic dehydrogenation units that can produce light olefins, in
particular, propylene and butylene, at high selectivities. This technology is known
as the UOP OleflexTM process.
The UOP Oleflex process can dehydrogenate propane, isobutane, normal-butane,
or isopentane feedstocks separately or as a mixture of spanning two consecutive
carbon numbers. The UOP Oleflex process and uses of the Oleflex process to
produce propylene, MTBE, and alkylate are discussed in Sections 3.3.2 and 3.3.3 of
this chapter.
As shown in Figure 3.3, the UOP Oleflex process is separated into three different
sections:
● Reactor section,
● Product recovery section, and
● Catalyst regeneration section.
Catalyst flow
C
C
R
Dryer
Rx effluent
compressor
Heater cells
Cold box
To frac
section
H2 Recycle
Fresh and
recycle feed Net gas (H2)
the reactors. The catalyst is collected on the bottom of the last reactor, transported
pneumatically to the regenerator. The effluent leaves the last reactor, exchanges heat
with the combined feed, and is sent to the product recovery section.
The slowly moving bed of catalyst circulates in a loop through the reactors and the
regenerator. The cycle time around the loop can be adjusted within broad limits but
is typically anywhere from 5 to 10 days. The catalyst activity can be restored with-
out interrupting the catalytic dehydrogenation process in the reactor and recovery
sections.
3.3 Light Paraffin Dehydrogenation 51
Collector
Disengaging
hopper Lift gas
blower
R
Regeneration
tower
R Flow control
Catalyst
lift lines hopper
Lock
hopper Lock
hopper Surge hopper
Lift engager
Lift
engager
Nitrogen H2
lift gas lift gas
The Oleflex process for light paraffin dehydrogenation uses a platinum catalyst
supported on alumina base. The key role of the dehydrogenation catalysts is to accel-
erate the main reaction while controlling other reactions. The alumina-supported
platinum catalysts are highly active but are not selective to dehydrogenation. In addi-
tion, the catalyst rapidly deactivates due to fouling by heavy carbonaceous materials.
Therefore, the properties of platinum and the alumina support need to be modified
to suppress the formation of byproducts and to increase catalyst stability. Arsenic,
tin, germanium, lead, and bismuth are among metals reported as platinum activity
modifiers [6]. Platinum is a highly active catalytic element and is not required in
large quantities to catalyze the reactions when it is dispersed on a high surface area
support. The high dispersion is also necessary to achieve high selectivity to dehy-
drogenation relative to undesirable side reactions, such as cracking. A typical high
surface area alumina supports employed having acidic sites that accelerate skeletal
isomerization, cracking, oligomerization, and polymerization of olefinic material,
and enhance “coke” formation. Alkali or alkaline earth metals assist in the con-
trol of the acidity [6]. Modified catalysts possess high activity and high selectivity
to mono-olefins. The major byproducts are controlled kinetically.
700
650
600
Temperature, ° C
550
450
400
300
2 3 4 5 6 7 8 9 10 11 12 13 14 15
Carbon number
Figure 3.5 Temperatures required to achieve 10% and 40% conversion of C2 –C15
normal-paraffins at 1 bar. Source: Bhasin et al. [6]. © 2001, Elsevier.
C3 H8 → C2 H4 + CH4
C2 H4 + H2 ↔ C2 H6
The main propane dehydrogenation is operated at equilibrium favored by high tem-
perature and low pressure. Figure 3.5 [6] shows the temperatures required for the
dehydrogenation of light paraffins are much higher than those for heavy paraffins.
For 40% conversion of propane requires a temperature of 600 ∘ C. It is advisable to
increase the conversion per pass as much as possible so that the cost of recycling
the unconverted reactant and that of separating the reactant and product is reduced.
However, high temperatures also enhance side reactions, such as coke formation,
causing fast catalyst deactivation and thermal cracking reactions that reduce the
selectivity to the desired product. Thus, all factors need to take into consideration
to optimize the process economics.
Oleflex process units typically operate in conjunction with fractionators and other
process units within a production complex. In a propylene complex (Figure 3.6),
a preheated propane-rich liquefied petroleum gas (LPG) feedstock is introduced
to a depropanizer to reject butanes and heavier hydrocarbons. The depropanizer
overhead is then directed to the Oleflex unit. The once-through propane conversion
closely approaches the equilibrium value defined by the Oleflex process conditions,
to selectively produce propylene and hydrogen with minimized side products. Two
product streams are created within the C3 Oleflex unit: a hydrogen-rich vapor
product and a liquid product rich in propane and propylene. The net hydrogen can
be exported directly, upgraded to PSA hydrogen, or used as fuel within the plant if
hydrogen is not in demand in the vicinity of the C3 Oleflex complex.
Trace levels of methyl acetylene and propadiene are removed from the Oleflex
liquid product by selective hydrogenation. The SHP process, consisting of a single
liquid-phase reactor, selectively saturates diolefins and acetylenes to mono-olefins
without saturating propylene.
3.3 Light Paraffin Dehydrogenation 53
Propylene
Oleflex
unit SHP
Deethanizer
Depropanizer
C3 LPG
P–P Splitter
C4+
Ethane and lighter material enter the propylene complex in the feed and are
also created by nonselective reactions within the Oleflex unit. These light ends
are rejected from the complex by a deethanizer column. The deethanizer bot-
toms are then directed to a propane–propylene (P–P) splitter. The splitter produces
high-purity propylene as the overhead product. Typical propylene purity ranges from
99.5 and 99.8 wt%. Unconverted propane from the Oleflex unit concentrates in the
splitter bottoms and is returned to the depropanizer for recycle to the Oleflex unit.
UOP’s latest Oleflex complex design with new generation of catalyst, DeH-26, con-
sumes less than 1.12 MT of propane per MT of propylene produced [7] (Figure 3.7).
Investment in a C3 Oleflex complex can be an excellent feedstock and end-product
diversification play to increase the operating flexibility of the refinery. The propane
feedstock from natural gas liquid (NGL) plants can be supplemented with propane
produced within the refinery and sent to the C3 Oleflex complex. The C3 Ole-
flex complex can also be used to upgrade lower-purity refinery-grade propylene
(RGP) streams produced by the FCC unit to chemical-grade propylene (CGP) or
polymer-grade propylene (PGP) in the propylene complex P–P splitter to realize
an additional product value uplift. Additionally, the hydrogen produced within the
Oleflex 1990
100% DeH-6 Catalyst
1.25 MTC3/MTC3=
Oleflex 2004
CCOP reduction from base
High
95% DeH-14
Performance
1.18 MTC3/MTC3=
Oleflex 2015
Oleflex 1997 Energy reduction Next generation
DeH-16
DeH-12
90% 1.14 MTC3/MTC3= Oleflex 2018
1.21 MTC3/MTC3= DeH-26
Energy
reduction <1.12 MTC3/MTC3=
Oleflex 2011 Further energy
85% DeH-16 reduction
1.16 MTC3/MTC3=
Energy reduction
80%
0 5 10 15 20 25 30 35 40
% ROI improvement from base
Propane PGP
RGP C3 Oleflex
Refinery
complex
H2 CGP
C3 Oleflex complex can be used for hydroprocessing needs within the refinery or
exported for sale (Figure 3.8).
Field butane, a mixture of normal-butane and isobutane obtained from natural gas
condensate, is fed to a deisobutanizer (DIB) column. The DIB column prepares an
isobutane overhead product, rejects any pentane or heavier material in the DIB bot-
toms, and makes a normal-butane side cut for feed to the butane isomerization unit.
H2 C3–
Methanol
Deisobutanizer
Field
butanes
Butamer
C5+
The DIB overhead is directed to the iso-C4 Oleflex unit, in which isobutane is selec-
tively converted to isobutylene and hydrogen. The main reaction and side reactions
are shown below:
Main reaction:
i-C4 H10 ↔ i-C4 H8 + H2
Side reactions:
i-C4 H10 ↔ n-C4 H10
n-C4 H8 ↔ C4 H6 + H2
Cracking of butanes to lighter materials also occurs as side reactions. The main
isobutane dehydrogenation is operated at equilibrium favored by high temperature
and low pressure.
Two product streams are created within the iso-C4 Oleflex unit: a hydrogen-rich
vapor and a liquid product rich in isobutane and isobutylene. The iso-C4 Oleflex
liquid product is sent to an etherification unit, where methanol reacts with isobuty-
lene to make MTBE. Raffinate from the etherification unit is depropanized to remove
propane and lighter materials. The depropanizer bottoms are then dried, saturated,
and returned to the DIB column (Figure 3.9).
Although MTBE is a good gasoline-blending component, it has properties that
cause problems. MTBE is more soluble in water than most other components of
gasoline. If the gasoline containing MTBE is spilled on the ground or leaks out of
underground storage tanks, MTBE could contaminate groundwater. Even fairly
small amount of MTBE in water can give it an unpleasant taste and odor, making
the water undrinkable. MTBE also does not break down (biodegrade) easily, so it is
harder to clean up once contamination occurs. In the late 1990s, many community
drinking water suppliers in the U.S. were found to have detectable levels of MTBE.
As a result, MTBE has been phased out for gasoline blending in the United States
in late 1990s. U.S. refineries were then faced with the challenge of replacing the lost
volume and octane value of MTBE in the gasoline pool. Isooctane was identified
as a cost-effective alternative to MTBE. It utilized the same isobutylene feeds used
in MTBE production and offered excellent gasoline-blending value. Furthermore,
isooctane production can be achieved in a low-cost revamp of existing MTBE plant
to the UOP InAlkTM process unit.
As shown in Figure 3.10, the Oleflex dehydrogenation unit can be easily inte-
grated with downstream dimerization of isobutylene followed by hydrogenation
(UOP InAlk process). The hydrogen from the Oleflex can be used to hydrogenate
isooctene to isooctane, a high-octane gasoline-blending component. The UOP
process for this combination is the UOP InAlk complex.
56 3 Olefins and Synthesis Gas Production Technologies
i-C4 Dimerization
Oleflex
InAlk
H2
Deisobutanizer
Hydrogenation
Field
butanes
Iso-octane
Frac section
Butamer
C3–
C5+
H2 C3–
Propylene
C3/i-C4 C3/C4
Oleflex separation
Methanol
Deisobutanizer
LPG Ethermax
Butamer MTBE
Frac
Section
C5+
H2
Butamer
dehydrogenation. UOP mixed-C4 Oleflex complex (Figure 3.12) has the following
features:
● Capability to process any mixture of butanes, including 100% normal-butane,
● Field butanes fed directly to the mixed-butane Oleflex – elimination of the need
for deisobutanizer, and
● Alkylate octane increased by 1–2 numbers ((R+M)/2) using the feed from
mixed-butane Oleflex relative to the use of 100% isobutylene feed.
Hydrogen-rich
off-gas
Light
Hydrogen C end gas
recycle gas
Reactor
Light end
Separator liquid
Charge Stripper
heater
Benzene
Kerosene Benzene
Light
Prefractionation ends Heavy
hydrotreating Aromatics alkylate
Hydrogen
Detal or
Pacol detergent
Purification PEP LAB
Define alkylate
Raffinate
return to Recycle Paraffin
refinery
Propylene
Isobutylene
Benzene
Ethanol conversion SAPO 34 exhibited the following features:
Regen gas
MTO process integrated with
Olefin cracking process (OCP)
C 4+
Air Water
OCP
Sep
Methanol section
By-products
100
Light Olefin carbon yields, wt%
90
80
70
1.0 1.2 1.4 1.6 1.8 20
Propylene/ethylene (P/E) product ratio, wt.
recovered and returned to the reactor. The reactor provides very high conversion so
there is no need for a large recycle stream. After the oxygenate recovery section, the
effluent is further processed in the fractionation and purification section to remove
contaminants and separate the key products from the byproduct components. Ethy-
lene and propylene are produced as polymer-grade products and sent to storage. The
C4 –C6 fraction is sent to the OCP reactor where it is selectively converted to light
olefins, majority of which is propylene. Typically, the propylene-to-ethylene ratio in
OCP reactor effluent is about 4. The OCP product is depropanized, the C3 and lighter
fraction is sent to MTO product recovery section, and residual C4 plus fraction is
taken as byproduct fuel.
As shown in Figure 3.17, the advanced MTO process, which is the integrated
MTO-OCP processes, can produce propylene to ethylene product ratio between
1.2 and 1.8 [16] to help meet the growing demand for propylene and additional
flexibility is achievable with the technologies, if desired.
Pipeline spec:
Acid gas
Gas • CO2 < 2 – 8%
CO2
H2O N2 Hg C2+ • H2S < 4 ppm
composition H2S
• Hg < 0.01 μg/Nm3
• H2O < 2 – 8%
CH4, C2+,
H2O, H2S, Pipeline gas
Raw CO2, N2, Hg
natural
gas
LNG
LNG spec:
Acid gas N2 Hg C2+ • CO2 < 50 ppm
H2O
CO2 • H2S < 2 – 4 ppm
H2S • Hg < 0.01 μg/Nm3
• H2O < 0.1 ppmv
Acid gas
Sweet Acid gas
gas KO drum
Make-up
water Acid gas
cooler
Amine Lean Lean
absorber solution solution Reflux
Flash gas pump cooler pump
to fuel
header Filtration
system Amine Make-up
Feed Make-up stripper water
gas water
Rich
flash drum
Amine
reboiler
Lean/rich Lean solution
exchanger booster pump
Figure 3.19 UOP amine guard FS process (1-stage) for acid gas removal. Source: Based on
“UOP Amine GuardTM FS Technology for Acid Gas Removal”, https://honeywelluop
.chinacloudsites.cn/wp-content/uploads/2011/02/UOP-Amine-Guard-Technology-for-Acid-
Gas-Removal-tech-presentation.pdf [17].
Benfield
absorber Condensate
pump
Lean
solution
filter
Feed
gas Hydraulic
turbine
(options)
Lean
Carbonate Condensate
solution
reboiler reboiler
flash
Lean solution
pump
Treated
gas
CO2
absorber
H2S Acid
CO2
Lean stripper gas
solution Reflux
filter accumulator
Sulfur Makeup
absorber H2S Reflux water
pump Export
Feed concentrator
gas water
Stripper
Packinox reboiler
exchanger
Figure 3.21 UOP selexol technology for H2 S removal and CO2 capture. Source: Based on
“UOP SelexolTM Technology Applications for CO2 Capture”, https://www.uwyo.edu/eori/_
files/co2conference09/mike%20clark%20revised%20selexol_wyoming_conference.pdf [19].
Permeate
(High acid
gas)
Permeate
Stage 1
Liquid
H2O
Feed gas
Liquid
Hg
Adsorption Adsorption Regeneration
Figure 3.24 UOP mercury removal and recovery system. Source: Based on Corvini et al.
[21].
needed for the subsequent gasification reactions. The remaining carbon reacts with
carbon dioxide, steam, and hydrogen to produce carbon monoxide, hydrogen, and
methane. Water gas shift and methanation are reversible reactions taking place
simultaneously based on the gasifier conditions.
There are three major types of gasifiers: moving-bed (also called fixed-bed),
fluidized-bed, and entrained-flow. Each type is defined by how the reactor brings the
coal in contact with the reacting gas and the temperature profile inside the gasifier.
Moving-bed reactor: In a moving-bed reactor, large particles of coal move slowly
down through the gasifier while reacting with gases moving up (Figure 3.25) [32]
at moderate pressures (25–30 bars). Reactions within the gasifier occur in differ-
ent “zones.” In the “drying zone” at top of the gasifier, the entering coal is heated
and dried, while cooling the product gas before it leaves the reactor. The coal is fur-
ther heated and devolatized by the high-temperature gas as it descends through
the “carbonization zone.” In the next zone, the “gasification zone,” the devola-
tized coal is gasified by reaction with steam and carbon dioxide. Near the bottom
of the gasifier, in the “combustion zone,” which operates at the highest tempera-
ture, oxygen reacts with the remaining char. Due to the countercurrent flow and
the inherent recuperation of the sensible energy in the gas through devolatiliza-
tion and coal drying, this type of gasifier gives the highest cold gas efficiency of
any of the gasifiers.
Fluidized-bed reactor: Fluidized-bed reactors employ back-mixing and efficiently
mix feed coal particles with coal particles already undergoing gasification in the
reactor vessel. Coal of small particles (<6 mm) is supplied through the side of
the reactor, and oxidant and steam are supplied near the bottom (Figure 3.26)
[32]. Due to the thorough mixing within the gasifier, a constant temperature is
sustained in the reactor bed. The gasifiers normally operate at moderately high
temperature to achieve an acceptable carbon conversion (e.g. 90–95%) and to
decompose most of the tar, oils, phenols, and other liquid byproducts. However,
the operating temperatures are usually less than the ash fusion temperature so as
to avoid clinker formation and the possibility of de-fluidization of the bed. This,
in turn, means that fluidized-bed gasifiers are best suited to relatively reactive
coals, low-rank coals, and other fuels such as biomass [32].
Entrained-flow reactor: In entrained-flow systems, fine coal feed and oxidant (air
or oxygen) and/or steam are fed co-currently to the gasifier (Figure 3.27) [32].
Entrained-flow gasifiers operate at high temperature and pressure with turbulent
flow, which causes rapid feed conversion and allows high throughput. The
gasification reactions occur at a very high rate (typical residence time is on the
order of few seconds), with high carbon conversion efficiencies (98–99.5%).
Entrained-flow gasifiers have the ability to handle practically any coal feedstock
and produce a clean, tar-free syngas. Given the high operating temperatures, this
type of gasifier melts the coal ash into vitreous inert slag. Different systems may
use different coal feed systems (dry or water slurry) and heat recovery systems.
A comparison of moving-bed, fluidized-bed, and entrained-flow gasifiers is shown
in Table 3.3 [32].
70 3 Olefins and Synthesis Gas Production Technologies
g g
Coal Gasifier
bottom
Gas
Coal
Gas
Steam,
oxygen
or air
Ash
Gasifier
Steam, bottom
oxygen 0 250 500 750 1000 1250 1500
or air Ash Temperature - °C
Gas
Gasifier
bottom
Coal Gas
Coal
Steam,
oxygen
or air
Raw syngas should be purified and the H2 /CO ratio should be adjusted before it
is used as feed for downstream processing. Desired H2 /CO values for various syngas
applications are shown below:
● H 2 production: Pure H2
● Ammonia: H2 /CO = 3
● Methanol: H2 /CO = 2
Steam,
Coal Gasifier
oxygen
bottom
or air
Coal Steam,
oxygen
or air
Gas
Gas Slag
Gasifier
bottom
0 250 500 750 1000 1250 1500
Temperature - °C
Slag
References
1 Foster, J. (2013). Can Shale Gale Save the Naphtha Crackers? London:
Platts.Platts special report: petrochemicals
2 Graig, R.G. and Spence, D.C. (1986). Catalytic dehydrogenation of liquefied
petroleum gas by the Houdry Catofin and Catadiene processes. In: Hand-
book of Petroleum Refining Processes (ed. R.A. Meyers). New York (Chapter
4.1: McGraw-Hill.
3 Sanfilippo, D., Buonomo, F., Fusco, G., and Miracca, I. (1998). Paraffins activa-
tion through fluidized bed dehydrogenation: the answer to light olefins demand
increase. Studies in Surface Science and Catalysis 119: 919–924.
4 Iezzi, R. Bartolini, A. (1997). Process for dehydrogenating light paraffins in a flu-
idized bed reactor. US Patent 5,633,421.
5 Luckenbach, E.C. Zenz, F.A. Papa, Giovanni Bertolini, A. (1997). Fluidized bed
reactor and process for performing reactions therein. US Patent 5,656,243.
6 Bhasin, M.M., McCain, J.H., Vora, B.V. et al. (2001). Dehydrogenation and
oxydehydrogenation of paraffins to olefins. Applied Catalysis A: General 221:
397–419.
7 Shakur, M. (2018). New generation of OleflexTM technology. Presented in UOP
International Technology Conference, Beijing, China.
8 Vora, B.V. (2012). Development of dehydrogenation catalysts and processes. Top-
ics in Catalysis 55: 1297–1308.
9 Chang, Silvestri, Smith, R.L. Aromatization reactions. US Patents 3,894,103;
3,928,483.
10 Kaiser, S.W. Production of light olefins. US Patent 4,499,327; 1985.
11 Kaiser, S.W. and Arabian, J. (1985). The Sciences and Engineering 10: 361.
12 Lewis, J.M.O. (1998). Methanol to olefins process using silicoalumino-phosphate
molecular sieve catalysts. Catalysis (ed. J.W. Ward). Amsterdam: Elsevier.
13 Chen, J.Q., Bozzano, A., Glover, B. et al. (2005). Recent advancements in ethy-
lene and propylene production using the UOP/hydro MTO process. Catalysis
Today 106: 103–107.
References 73
4.1 Introduction
As discussed in Chapter 3, ethylene and propylene are the main petrochemical build-
ing blocks. Ethylene is used as a feedstock to manufacture polyethylene, ethylene
chloride, ethylene oxide (EO), and other petrochemicals (Figure 4.1). These mate-
rials are very useful for the packaging, plastic processing, construction, houseware,
and textile industries. Most propylene is used to make polypropylene, but it is also a
feedstock to produce propylene oxide, acrylic acid, cumene, and many other chemi-
cal derivatives (Figure 4.2). Not only plastic processing, the packaging industry, and
the furnishing section, but also the automotive industry are users of propylene and
its derivatives.
In addition to ethylene and propylene, other olefins also find uses in fuel and petro-
chemical chemical applications. As discussed in Chapter 3, the leading applications
of butylene are for production of gasoline-blending components (alkylate, MTBE,
and ETBE) while heavy normal-olefins are primarily used for producing linear alkyl-
benzene (LAB), a dominant precursor of biodegradable detergents. Syngas, on the
other hand, represents a suitable alternative, from resources other than crude oil, to
produce chemicals with commercial significance such as hydrogen, ammonia, and
methanol. In this chapter, the uses of olefins (ethylene and propylene) and syngas
as raw materials to produce downstream products will be discussed.
Polyethylene
Others
Propylene
Acrylic fibers, ABS
Acrylonitrile
resins, nylon, etc.
Others
Ethylene
Ethylene Pyrolysis Vinyl Polymerization Polyvinyl
dichloride chloride chloride
(EDC) (VCM) (PVC)
Chlorine
short chains in LLDPE are able to move against on another upon elongation without
entangling together, resulting in improved strength and impact properties.
contents of the reaction vessel are pressurized and continually mixed to maintain
the suspension and ensure a uniform particle size of the PVC resin. The reaction is
exothermic and thus requires cooling. As the volume is reduced during the reaction
(PVC is denser than VCM), water is continually added to the mixture to maintain
the suspension [5].
Air or high-purity oxygen is mixed with recycled gas and fresh ethylene [6].
Vapor-phase oxidation inhibitors such as EDC or vinyl chloride or other halo-
genated compounds are added in the inlet of the reactor to retard carbon dioxide
formation [7]. The gas is fed to s multi-tube reactor in which the reactor temperature
is controlled at 220–300 ∘ C and pressure is set at 10–30 bars. The oxygen-based EO
process is preferred over the air route in large plants due to higher yields and less
catalyst required [7].
EO is further processed to produce MEG by a liquid-phase, noncatalytic hydrolysis
reaction operated at an elevated temperature.
H2C CH2 + H2O HOCH2CH2OH
O
This conventional process was started in 1937 by Union Carbide [6]. In this reac-
tion system, MEG inevitably reacts with remaining EO generating diethylene glycol
and triethylene glycol.
Selectivity to the different glycols is controlled by varying the ratio of water to EO
with large excess of water promoting MEG selectivity. Removing the excess water is
energy intensive and requires capital investment in evaporators. These factors limit
the amount of excess water, which can be used for control of the selectivity to MEG.
To maximize the MEG selectivity, many different systems have been studied. Mit-
subishi and Shell developed a two-step process [8] called OMEGA process.
C2 H4 O + CO2 → C3 H4 O3
C3 H4 O3 + H2 O → HOC2 H4 OH + CO2
In the first step, EO is reacted with CO2 to form ethylene carbonate. In the second
step, the ethylene carbonate is catalytically hydrolyzed to MEG and CO2 released in
this second reaction step is recycled back to the first reaction step.
4.3 Uses of Propylene 79
Atactic polypropylene
Isotactic polypropylene
Syndiotactic polypropylene
There are two major processes for the manufacture of propylene oxide: propylene
chlorohydrin process and propylene oxidation process using peroxides.
The chlorohydrination process consists of formation of propylene chlorohydrin by
the reaction between hypochlorous acid and propylene. The propylene chlorohydrin
is epoxidized to propylene oxide by a 10% solution of lime or NaOH.
Cl2 + H2O HOCl + HCl
Tail
gas
References 83
to produce ammonia.
N2 + 3 H2 ↔ 2 NH3
A wide range of reaction conditions are used to produce ammonia, temperature
between 400 and 500 ∘ C and pressure between 150 and 250 bars. As the reaction is
exothermic, cool reactants (nitrogen and hydrogen) are added to maintain stable
temperatures of the reactors. On each pass, only about 15% conversion occurs, but
any unreacted gases are recycled, and eventually an overall conversion of 97% can
be achieved [16].
Improvements have been made in ammonia synthesis catalysts. Traditionally,
ammonia catalysts have been ion-based materials with promoters. Recent research
has shown a ruthenium-based catalyst is much more active to enable the process to
take place at lower pressures.
References
4 Chanda, M. and Roy, S.K. (2006). Plastics Technology Handbook, 1–6. CRC Press.
5 Allsopp, M.W. and Vianello, G. (2012). Poly(vinyl chloride). In: Ullmann’s Ency-
clopedia of Industrial Chemistry (eds. Barbara Elvers, Stephen Hawkins, and
William Russey), pp. 717–742. Weinheim: Wiley-VCH.
6 McKetta, J.J. (1984). Encyclopedia of Chemical Processing and Design, vol. 20.
New York: Marcel Dekker Inc.
7 Othmer, K. (2004). Encyclopedia of Chemical Technology, 4e, vol. 9. New York:
Marcel Dekker Inc.
8 Shell Magazine 2006. Spring; https://www.shell.com/business-customers/
catalysts-technologies/licensed-technologies/petrochemicals/ethylene-oxide-
production/omega-process.html.
9 Whiteley, K.S., Geoffrey Heggs, T., Koch, H. et al. (2005). Polyolefins. In: Ull-
mann’s Encyclopedia of Industrial Chemistry (ed. Claudia Ley), pp. 36–69.
Wiley-VCH: Weinheim.
10 Devesh, Tripathi (2002). Practical guide to polypropylene. UK: Shawbury.
https://www.worldcat.org/oclc/568032693
11 Gahleitner, M. and Paulik, C. (2014). Polypropylene. In: Ullmann’s Encyclopedia
of Industrial Chemistry, 1–44. Wiley-VCH Verlag GmbH & Co. KGaA. https://doi
.org/10.1002/14356007.021_o04.pub2. IBBN 9783527306732.
12 Nijhuis, T.A., Makkee, M.M., Moulijn, J.A., and Weckhuysen, B.M. (2006).
The production of propene oxide: catalytic processes and recent developments.
Industrial & Engineering Chemistry Research 45 (10): 3447.
13 Brazdil, J.F. Acylonitrile. In: Ullmann’s Encyclopedia of Industrial Chemistry (ed.
Claudia Ley), pp. 1–10. Weinheim: Wiley-VCH.
14 Ngoh, S.K. and Njomo, D. (2012). An overview of hydrogen gas production from
solar energy. Renewable and Sustainable Energy Reviews 16: 6782–6792.
15 Cho, W.C., Lee, D.Y., Seo, M.W. et al. (2014). Continuous operation character-
istics of chemical looping hydrogen production system. Applied Energy 113:
1667–1674.
16 Appl, M. (2006). Ammonia. In: Ullmann’s Encyclopedia of Industrial Chemistry
(ed. Claudia Ley), pp. 1–33. Weinheim: Wiley-VCH. https://doi.org/10.1002/
14356007.a02_143.pub2.
85
Section III
Aromatics
87
5.1 Introduction
Aromatics are cyclic hydrocarbons with delocalized π electrons forming a planar
ring system with alternating double and single bonds. With the resonance of the
conjugated bonds in a ring system, aromatics are stable molecules and yet highly
reactive with excellent ability for electrophilic and nucleophilic aromatic substitu-
tions. Historically, aromatics were originally found as a lighting gas in the eighteenth
century. Benzene was later discovered in 1825 by Michael Faraday by recovering
the condensed part of a lighting gas found in whale oil. Benzene in coal tar was
later found in 1845 by August Wilhelm von Hofmann. The early findings show that
coal tar possesses a variety of aromatic components that can be used as fuels. In
the 1920s, the primary sources of aromatics came from coal tar and coke-oven ben-
zole, which derive from thermal processing of coal. With excellent octane as a fuel
property, aromatics were in high demand for fuel supply during World War II. In
1940, Standard Oil Company and M.W. Kellogg Limited developed the hydroform-
ing process to upgrade virgin naphthas. Toluene was extracted to produce explosives
(trinitrotoluene). The process used molybdenum oxide on alumina as a catalyst. In
1949, UOP innovated and commercialized the first platinum-based reforming pro-
cess, the PlatformingTM process. It provided better yields and higher-octane products
for motor fuels. The catalytic reforming process also has contributed to the petro-
chemical industry as the primary source of industrial aromatics, overtaking conven-
tional coal tar and coke-oven production.
C5/C6
Isomerization
Naphtha splitter
Naphtha
feed Gasoline
NHT
pool
C7+
Reforming
N-Paraffins
M or A M/A
A M
Cracked Cyclopentanes Cyclohexanes Aromatics
products
M/A
M or A
Isoparaffins III
Naphthene Naphthene
isomerization dehydrogenation
I II III
Naphthene
R R Figure 5.3 Major reactions happening in the
dehydrogenation S + 3H2
reforming reactor.
C
Paraffin R C C C C R C C C
isomerization
R Rʹ
Naphthene S
isomerization
S
Rʹ
S + H2
Paraffin
R C C C C
dehydrocyclization
Rʹ
S + H2
C C
Hydrocracking R C C C + H2 RH + C C C
H
R Rʹ
Aromatics H2 Rʹʹ
+ +
dealkylation
140+ Units 100+ Units 115+ Units 25+ Units 4 Units 15 Units 10+ Units
Low Coke
Shape Sphere
Pill diameter 1/16 in. (1.6 mm)
Surface area 180 m2 /g
Density (ABD) 35 lb/ft3 (560 kg/m3 )
Pt level 0.29 wt%
Crush strength 50+ N
Promoter No Yes No
excellent surface area stability and chloride retention as the R-130 series. R-264
catalyst was introduced to markets that offer high yield, so customers can increase
their throughput. With higher density, this catalyst has increased reactor pinning
margin. With high-throughput benefit, the refiners can dial up their feed rate, and
with higher activity, maximum BTX aromatics production can be achieved. R-254
and R-284 are promoted catalysts that can offer higher yields. UOP introduced 300
series products with R-334 catalyst in 2013 and the latest R-364 catalyst in 2017 pro-
viding highest yield, long life, and low coke. These catalysts, however, do not include
a promoter and are primarily optimized for high performance by using proprietary
base and manufacturing technique (Table 5.1). The latest-generation R-364 builds
on the success of UOP’s proven R-264 catalyst with over 100 loads sold. R-364
provides great flexibility in operation for the refiner with two different modes, high
activity or high yield.
5.2 Aromatic Production 93
The first catalytic reforming unit was installed in 1949. The unit was the
first commercial UOP designed semi-regenerative process, which includes three
fixed-bed reactors. The use of a noble metal catalyst has enabled catalytic reforming
a workhorse to produce high-octane gasoline. Typically, the semi-regenerative
reforming unit can operate from 12 to 24 months until the catalyst loses its activity
and selectivity. Catalyst is then regenerated in situ. Thus, higher-pressure operation
generally favors cycle length, keeping the unit onstream longer.
The catalytic process has evolved over times from semi-regenerative, cyclic
reforming to continuous catalytic regeneration (https://www.uop.com/processing-
solutions/refining/gasoline/ https://www.uop.com/processing-solutions/refining/
gasoline/#naphtha-reforming).
Process Variables:
● Reactor temperature
● Reactor pressure
● Space velocity
● Hydrogen/hydrocarbon ratio (H2 /HC ratio)
● Feedstock
Reactor Temperature
This is an important variable that directly controls the severity of the process to
affect activity, overall yields, or product quality (RON). High severity is referred
to as high-temperature operation while keeping other conditions constant. At high
severity, high-octane reformate can be achieved at the expense of catalyst activity.
Typical range for reactor inlet temperature is 490–550 ∘ C. Higher-range temperature
increases the rate of catalyst deactivated with rapid coke laydown leading to more
frequent regeneration.
Reactor Pressure
A reforming reactor can typically run at 3.5–30 bar pressure. Lower pressures favor
higher reformate and H2 yield, while higher pressures increase longer life of cata-
lyst with lesser coke buildup with lower reformate yield compared to lower-pressure
operation. The lowest operating pressures are only run in reforming units with CCR.
Note that high-pressure operations tend to give lower reformate yield and more light
end byproducts. This is a direct result of undesirable hydrocracking of paraffin com-
ponents favoring light end formation rather than aromatization or ring formation.
Space Velocity
Space velocity is defined as the ratio of the volumetric feed rate to the amount of cat-
alyst. A common acronym for space velocity is liquid hourly space velocity (LHSV)
in units of h−1 . It measures how fast the feed passes the catalyst. Lower space velocity
means larger catalyst volumes (more active sites available), and severity or temper-
ature can be lower to achieve the same activity or product quality. However, higher
space velocity (by pushing more feed flow) requires higher temperature to increase
reaction rates for the same product target. With fixed design of the reactor for a given
capacity, LHSV can only be varied by changing the naphtha charge rate to the unit.
94 5 Aromatic Production Technologies
Feedstock
Feedstocks to a reforming unit are characterized by distillate cut points and com-
ponents, including paraffins, naphthenes, and aromatics (PNA). Feeds with high
paraffin components are harder to convert to aromatics due to slow reaction rates.
Naphthene components convert at a much faster rate and get almost consumed in
the very first reactor. Thus, rich naphthene feedstocks would require less severity to
reach a product octane target and aromatic concentration. On the other hand, lean
naphthene feedstocks (more paraffin concentration) require more severe operating
conditions to achieve the product quality target. Regarding boiling range, feedstocks
with a higher end point or a wide boiling range require more severe operating con-
ditions. If the target product is fed to an aromatic complex, the boiling range of the
feedstock to the reformer is limited to a more narrow cut range so that the desired
aromatics, such as xylenes (and less C9 aromatics), can be obtained.
CCR
platforming
C5+yield
Advantage
Fixed-bed
platforming
Pressure
Figure 5.6 Reformate yield-pressure relationship for CCR and fixed-bed platforming.
UOP CCR Platforming is the leading catalytic technology with more than 250
units licensed in the world, producing rich aromatic reformates for either refining
or petrochemical uses. UOP RZ Platforming is another technology that converts
C6 and C7 materials in the raffinate stream rejected from extractive distilla-
tion to produce additional aromatics. With RZ-100 catalyst used in a fixed-bed
reactor system, RZ Platforming offers a way to convert difficult components to
aromatics, thus reducing the naphtha feed requirements for a target aromatics
capacity.
Net gas
Fuel gas
Rx Rx Rx C S
F E
E P
LPG
H H H D
E
B
Feed H
Legend
Rx Reactor Reformate
H Heater
CFE Combined feed exchanger
SEP Separator
DEB Debutanizer
(a)
Flue gas
H2-rich gas
Swing
Reactor Reactor Reactor
reactor
Overhead
Air
Inert gas Furnace
Separator
Pretreated
naphtha
Furnace Furnace Furnace
Reformate
(b)
Stacked
reactor
Naphtha feed
from Net H2 rich gas
Rx Net gas
treating
CCR compressor Fuel gas
regenerator Rx Recovery
section
Rx
Light ends
C S Stabilizer
Rx
F E
E P
H H H H
Fired heaters
H Aromatics
Spent catalyst
rich reformate
Legend
CFE Combined feed exchanger
SEP Separator
(c)
90 Theoretical yield
C5 yied, Lv%
1990s
86 1970s 1980s
1960s
1950s
82
78
86 90 94 98 102 106
RON clear
conditions are reported at 350–530 ∘ C, and 20–50 bar [4]. The xylene products
obtained are in equilibrium with 23–25% p-xylene, 50–55% meta-xylene, and
23–25% ortho-xylene. Also, note that the equilibrium xylenes produced in the
Tatoray process contain low ethyl benzene (EB). This low EB concentration product
makes Tatoray process attractive and ready for isomer separation using ParexTM
unit or para-xylene crystallization unit.
For reaction mechanism, the two main pathways are toluene disproportionation
and transalkylation of toluene and C9 aromatics. With methyl group ability to
jump over from one ring molecule to another in the presence of the catalyst, the
methylated aromatic products are limited by equilibrium. Figure 5.10 illustrates the
equilibrium compositions of aromatic products for any given feed methyl/phenyl
ratio. For 100% feed, the methyl to phenyl ratio is 1, while all C9 aromatic feed would
have a 2 : 1 of methyl to phenyl ratio. The more methyl groups available in the feed,
the more chances or favorable equilibrium to proceed toward alkylated aromatic
products. However, two high methyl/to methyl ratios also lead to production of
heavier alkyl aromatics such as C10 and C11 aromatics. Thus, a typical 50 : 50
98 5 Aromatic Production Technologies
100%
Benzene
90%
80%
Composition, (wt%)
70%
Toluene
60%
50%
40%
30%
C8 aromatics
20%
C10 aromatics
10% C9 aromatics
0%
1.00 1.17 1.30 1.40 1.49 1.55 1.63 1.72 1.84 2.01 2.28
Feed methyl/phenyl ratio
Figure 5.9 Equilibrium compositions of aromatics with various feed methyl/phenyl ratios
at 700 K.
Toluene from
toluene column
Cg aromatics
from Ag column
Overhead liquid
Toluene from
Parex unit
Recycle gas
Product
mixture feed of toluene and C9 aromatics is a good feed to produce xylenes. Con-
version per pass is from 46% to 59% [4]. Higher conversion reduces the amount of
recycle materials through the benzene-toluene fractionation section of the complex
for equipment cost and utility consumption considerations. However, pushing
high conversion may result in benzene and xylene selectivity drop, faster catalyst
deactivation, and increase rates of ring loss or formation of heavy aromatic by
products.
5.2 Aromatic Production 99
Disproportionation
2 +
Toluene Benzene Xylenes
Transalkylation
+
Toluene C9 aromatics Xylenes
Raffinate
Sulfolane Benzene
PlatformingTM
B T
TatorayTM
NHT Unit
A11+
p-Xylene
Naphtha
ParexTM IsomarTM Light ends
unit unit
C2H5
TM
I-300 catalyst H2
EB dealkylation + C2H4 + C2H6
The Isomar flow diagram is shown in Figure 5.12. The feed usually comes from
the Parex unit or referred to as Parex raffinate with depleted para-xylene. The feed
is then combined with hydrogen-rich recycle gas and make-up gas. Small amount
of hydrogen is consumed across Isomar reactor. The combined feed is heated and
vaporized to reactor operating temperature. The reactor is a fixed-bed radial-flow
reactor where hot feed is passed radially through the catalyst bed. The reactor efflu-
ent stream is heat exchanged with the feed and sent to the product separator where
liquid and gas are recovered. A small gas stream is purged to avoid light ends buildup
in the recycle gas. The liquid effluent is stabilized in the deheptanizer where over-
head gas is taken as fuel gas to supply fuel in the complex. The overhead liquid is
usually sent back to sulfolane unit for benzene recovery. The deheptanizer bottoms
product is sent to xylene splitters for separation and eventually makes its way back
to the Parex and Isomar units.
102 5 Aromatic Production Technologies
Parex
raffinate
Fuel Gas
Charge
heater Purge
Reactor gas
Separator
To benzene
recovery
Deheptanizer
Steam
Makeup
hydrogen To xylene splitter
100
EB dealkylation EB isomerization
Product yield, (wt%)
90 6.4 0.2
80 86.3 87.6
70
60
50
I-350 I-400
PX Benzene
given xylene column and isomerization unit, Parex process can deliver more than
50% more p-xylene with higher recovery of separation per pass.
Desired reaction:
CH3
Side reactions:
+ CH3OH
+
Acid
Sequential methylation: 2 CH 3OH CH3 CH3
–2H2O –H2O
CH3 CH3 CH3
Transalkylation:
CH3 CH3 CH3
Acid Acid
+ +
CH3
CH3
1
R
Sulfolane Raffinate
TatorayTM
unit Heavy aromatics
Toluene
methylation
Methanol
A11+
TM
Platforming p-Xylene
ParexTM IsomarTM Light
unit unit ends
NHT
Naphtha
and utility consumption. The same can be said for existing units using crystallizer
for para-xylene separation instead of adsorptive separation. For existing aromatic
complexes, more than 20% capacity of para-xylene can be achieved using toluene
methylation with little changes in the process units. Toluene methylation is thus a
viable option for on-purpose para-xylene production particularly for regions with
cheap and abundant feedstocks such as China and Middle East.
Reactor Regeneration
section section
Tail
Hydrogen gas
Gas recovery
C
C
R
Light ends
Heater cells
Fresh and
recycle feed
Stripper
Fresh
LPG feed
C6+
aromatics
5.2.5.1 Mechanism
The reaction mechanism for LPG and condensate aromatization process is driven by
dehydrocyclodimerization chemistry, which favors at 425–500 ∘ C temperature. The
reaction pathway includes the following steps:
and ring formation. This can be managed by coke burns and catalyst regeneration
in the CCR system.
Light
paraffins Olifins Oligomers Naphthenes Aromatics
The aromatics yield increases with the carbon number in the original feedstock
(butane vs. propane). Low-pressure operation with high temperature also favors
high yield of aromatics. The liquid product contains greater than 90% BTX aromatics
depending on the feed composition. The overall aromatics yield for propane or
butane-rich feed can be 60–67%. Again, butane-rich feed would produce higher
overall aromatics yield. Hydrogen is a significant byproduct and up to 5 wt% of the
feed can be recovered as high-purity H2 with standard PSA technology
Also, with design flexibility, Cyclar unit can be configured or revamped to turn to
an olefin production unit such as propane dehydrogenation unit (PDH or UOP Ole-
flex) or repurposed to a naphtha-reforming unit. Starting from a small-chain-length
paraffin (C3–C4 in LPG), the molecule dehydrogenates to an olefin, which has more
active functional group that can then oligomerize to a dimer olefin. The ring clo-
sure can result in an aromatic molecule and at the same time produce hydrogen as
a byproduct. The aromatization of the dimer molecule is an acid-catalyzed reaction.
The commonly used catalyst is ZSM-5 as a base with a promoted metal.
2-Ring aromatics
hal
pht
Higher carbon #
Na
1-Ring aromatics
lefins
enes, o
naphth
r p a raffins,
ola
Non-p
high aromatic content (usually means good octane) makes LCO a low cetane value
stream, with typical cetane from 15 to 25. In the US, significant amounts of LCO can
be blended to the diesel pool in up to 20% with similar boiling range with diesel – 95%
cut point at 360 ∘ C. LCO also has high sulfur and nitrogen content, up to 1.5 wt%
and 750 wt ppm, respectively. Note that major aromatic components are multiring
aromatics, 2-ring, and 3-ring components, shown by quantitative two-dimensional
GC × GC analysis. For aromatics production, these multiring components need to
be selectively converted or ring opened to single-ring aromatics. This is a suitable
strategy for the current trend of highly integrated refinery and petrochemical com-
plexes that can operate their FCC to maximize not only propylene but also aromatics
to maximize profits.
LCO upgrading mainly includes hydroprocessing for contaminant removals,
cetane improvement, or cut-point adjustments to meet fuel specifications.
Hydrotreating
(diesel, fuel oil)
LCO
Full conversion hydrocracking
(full range naphtha)
LCO UnicrackingTM process
(diesel, gasoline)
LCO-X process
(BTX, Lt. naphtha, LPG)
For a typical 60 000-barrel-per-day (bpd) FCC feed capacity, LCO yield can
be obtained at 14 000 bpd and sent to LCO-X process, which can produce
215 000 tons/year of mixed xylenes and 80 000 tons/year of benzene as petrochemical
5.2 Aromatic Production 109
LCO
feed
HT Light
LPG naphtha
HC
Aromatics Benzene
Off
maximization
gas
Mixed
xylenes
ULSD
blend
stock
products. Other byproducts include fuel gas, LPG, light naphtha, and ultra-low
sulfur diesel (ULSD). The amount of xylenes produced can easily be picked up by the
demand growth of para-xylene. For economy-of-scale considerations, LCO inter-
mediates from other FCCs can be routed together to a larger LCO-X unit to produce
more xylenes to gain healthy margin profits and quick return on investment.
A general process diagram for UOP LCO-X process is shown in Figure 5.17. The
reactor includes a multizone hydrotreating combined with hydrocracking catalysts.
The reactor effluent is sent to a hot high-pressure separator where recycle gas is
taken off in the overhead and returned to the reactor and the hot liquid is sent to
a low-pressure separator. Offgas is separated out and the separator liquid is routed
to fractionation section for product recovery. Several products can be obtained from
LCO-X process, including fuel gas, LPG, light naphtha, aromatics, and ultra-low
sulfur diesel.
From process chemistry point of view, the catalyst is designed to convert multiring
aromatics to single-ring components. The simplified reaction pathway is depicted in
Figure 5.18. Di-alkyl naphthalene is desired to cleave to 2 monoaromatic molecules.
Some saturation of the ring, whether complete or partial saturation, can occur at a
certain extent to accommodate the subsequent ring opening or dealkylation. Thus,
some naphtha range materials may be formed as intermediate products and aromatic
maximization reactor can further convert the alkyl-cyclohexanes to corresponding
aromatics (Figure 5.18).
110 5 Aromatic Production Technologies
R R
R
Aromatics
R maximization
reaction
ed
R sir n R
De ctio
rea
R Hy R
dro R R
ge
na R
tio
n
R
Naphtha range
LCO hydrocarbon distribution by cuts Coker gas oil hydrocarbon distribution by cuts
% Hydrocarbon type
% Hydrocarbon type
100% 100%
3 Ring
80% aromatics 80%
2 Ring
60% 60%
aromatics
40% 1 Ring aromatics 40%
20% 20%
0% 0%
Boiling point Boiling point
Product quality
Benzene ● 90–99.9%
Mixed xylenes ● Suitable for Parex unit feed for p-xylene
● Exceeds ASTM D-5211 specification
● ∼1% ethylbenzene
Light naphtha ● 80–90% C5 –C7 paraffins, 10–20% naphthenes
● 76–82 RON, high Iso content
LPG ● <0.5% olefin
● 60% C4 , 40% C3
Diesel ● <10 ppm sulfur, Ultra-Low-Sulfur Diesel (ULSD)
● 10–15 cetane increase over LCO
Refiners can re-configure or integrate their FCC to export aromatics for petro-
chemical production. BTX can be extracted using extractive distillation processes
with commercial proprietary solvents such as Carom, Udex, or Sulfolane (developed
by UOP and Shell).
UOP licensed LCO-X technology levering UnicrackingTM technology to convert
the byproduct LCO from FCC to convert multiring aromatics to single aromatics.
5.2 Aromatic Production 111
Heavy CCR
naphtha Platforming
Aromatics para-Xylene
complex benzene
Feed;
ethane
propane
butane
naphtha
gas oil Oil and
Cracking Cracked gas precooling C2
water Deethanizer Demethanizer C2 splitter
furnaces compression drying hydrogenation
quench
Pyrolysis gasoline
Propane recycle
C5+
Steam-cracking Yields
Steam crackers are a major workhorse to produce ethylene, a critical monomer for
the production polyethylene and other polymer products. Other olefins such as
propylene and butenes are byproducts. Table 5.2 summarizes typical yields from
steam cracking from different feedstocks with the basis on the same amount of
ethylene production. Pyrolysis gasoline includes high-octane aromatic components
such as benzene, toluene, and xylene. Pyrolysis gasoline yields increase with heavier
hydrocarbon feeds. However, due to the nature of thermal nonselective cracking,
5.2 Aromatic Production 113
Product Feedstock
Ethane Propane Butane Naphtha Gas oil
Composition %/ppm
Benzene 67%
Toluene 16%
pX, mX, oX, EB 6%
Styrene 1.3%
C9+ aromatics 7%
Nonaromatics 2%
Clean
coke oven gas
H2SO4
Coal Flushing
Coke ovens Liquor Ammonium
Benzol
sulfate
distillation
Tar distillation
plant
Tar product
Figure 5.21 Coke-oven flow diagram with coke-oven benzol recovery unit.
80
70
60
Relative yield (%)
50
Benzene Toluene
40
Xylenes
30
20
10
0
Reforming Pyrolysis Benzole
By-product
Biomass gas
Condenser
train
Feed
Reheater
handling
Reactor
Bio-oil
Blower Exhaust
gases
Blower
By-product
ash
as 70–75 wt%. Other products include char and light gas. The liquid product con-
tains high concentration of oxygenated hydrocarbons. Nonoxygenated hydrocarbon
products or BTX aromatics are also present but in much small quantities and can be
very difficult to recover. The liquid product obtained from biomass pyrolysis can be
referred to as pyrolysis oil, bio-oil, or bio-crude.
A UOP joint venture commercial pyrolysis process is shown in Figure 5.23. This
is a rapid thermal process referred as UOP RTPTM leveraging experiences from UOP
fluid catalytic cracking (FCC) technology. Biomass, in the absence of oxygen, is
pyrolyzed in a fluidized reactor at 500 ∘ C and quickly quenched to yield 65–75 wt%
RTP green fuel (Table 5.4). Utility is minimized by circulating hot sand inside the
fluidized reactor. Char and sand are disengaged in a cyclone and can be reheated
and cycled back to the reactor. Lighter char is lifted and disengaged in another
cyclone and discharged as byproduct ash.
The energy content per volume basis for RTP green fuel is about half of typical
fuel oil (No. 6) (Table 5.5). Thus, for a boiler or furnace requirement, about twice the
volume of RTP green oil is needed to substitute for conventional fuel oil. This can
be economical if the biomass feedstocks are readily abundant and cheap. However,
for transportation fuel market, the bio-oil needs to be further processed or refined
to lower the oxygen content while meeting the other fuel specifications required for
motor fuels.
CH 2OH CH 2OH
HC HC
HC HC
CH 2OH CH 2OH
CH 2OH CH 2OH
CH CH
CH CH CH
CH H3CO
H3CO
O CH O CH
CH 2OH
OCH 3 OCH 3
OH HC O
CHOH
OCH 3
OH
Tetralignol Pentalignol
Typical woody biomass contains about 50% cellulose, hemicellulose, and lignin.
Lignin component (as high as 25%) has high aromaticity and complex struc-
tures, which include multiple linkages of three basic monolignol building-block
molecules: p-coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol. These
monomers have a common phenolic backbone with alkyl- or oxygen groups
118 5 Aromatic Production Technologies
H2 O 17.9
C1–C5 25.2
o
C6–150 C 8.3
150–240 o C 5.7
150–240 o C phenols 37.5
240–260 o C 8.7
Residue 2.4
Hydrogen consumption 5.7
5.3 Summary
The aromatic production technologies continue to evolve. Catalytic reforming
provides the major production of aromatic building blocks. Industrial licensors
continue to develop catalysts with better performance and selectivity for aromatics
while minimizing byproducts such as light ends or LPG. At the same time, design
improvements, data analytics, and process intensification will help refineries utilize
their resources at the highest efficiency, increase profits/margins, and minimize
wastes and downtime.
On-purpose and secondary conversion processes continue to provide pathways to
fine-tune the distribution of products to high yields of desired products. To achieve
this, reaction and separation engineering will continue to be integrated to overcome
equilibrium limitation. Novel production methods taking advantage of in situ reac-
tive separation will emerge as a breakthrough challenging the current/traditional
technologies that have been used over the past 70 years. This will require new tech-
nologies or hybrid systems making use of new catalysts, adsorbents, and membrane
materials to improve the performance and yields while lowering costs and invest-
ment barriers.
References
Uses of Aromatics
6.1 Introduction
Aromatics make excellent monomers that can be crosslinked to synthesize valuable
polymers. With excellent substitution ability, high reactivity, good solvency, and
increased reactivity with substituent groups, aromatics continue to be the key
materials for polymer syntheses and development in material science. Essential
materials such as nylons, polyesters, plastics, pesticides, coatings, and resins are
primarily derived from basic aromatic building blocks, including benzene, toluene,
and xylene.
Benzene
Styrene Polystyrene
SBR elastomer
Ethylbenzene
Phenolic resin
Phenol Caprolactam
acetone Bisphenol A
Methyl methacrylate
Methyl isobutyl ketone
Cumene
Ethylbenzene 49
Cumene 21
Cyclohexane 12
Nitrobenzene 9
Detergent alkylation 3
Chlorobenzene 2
Maleic anhydride 2
Other 2
The world consumption of benzene in 2016 was 44 206 thousand metric tons and
summarized in Table 6.1 (IHS Markit [2]). It is projected to grow at 2.6% average
annual growth rate.
CH3
+ H2C CH2
CH3 CH2
+ H2
CH2 CH
Cross-linked polystyrene-divinylbenzene
CH2 CH
CH2
CH2 CH
CH2
CH2
n
Nylon 6, Nylon 6,6, PET, Polyurethane, Polystyrene, Polycarbonate…
Cyclohexane
Hexamethylene
Caprolactam
diamine
Caprolactam 55
Adipic acid 22
Cyclohexanol/cyclohexanone 20
Other 3
The total world consumption of cyclohexane in 2017 was 5223 thousand metric
ton. The major use is for nylon synthesis. The relative demand is summarized in
Table 6.2.
nH2N–(CH2)6–NH2 + nHOOC–(CH2)4–COOH [H2N–(CH2)6–NHCO–(CH2)4–CO]n
Hexamethylenediamine Adipic Acid Nylon 6,6
H
N
–H2O –H2O
C= O nH2N–(CH2)5–COOH [H2N–(CH2)5–CO]n
O OH
Co salt
+
HNO3Cu
Ammonium vandate
HOOC–(CH2)4–COOH
Adipic Acid
Or Benzene Phenol
6.2 Polymer Production from Basic Aromatic Building Blocks 125
Production of Hexamethylenediamine
DuPont Synthesis O O
2NH3 –H2O
HOOC (CH2)4 COOH H4NOOC (CH2)4 COONH4 H4NC (CH2)4 CNH4
H2 –2H2O
H2N (CH2)6 NH2 NC (CH2)4 CN
Hexamethylenediamine Adiponitrile
Cl2 NaCN
H 2C CH CH CH2 ClH2C CH CH CH2Cl NC CH2 CH CH CH2 CN
H2
H2N (CH2)6 NH2 NC (CH2)4 CN
Hexamethylenediamine Adiponitrile
HCN
H2C CH CH CH2 H3C CH CH CH2 CN
1,4-Addition
Butadiene Ni cat. with ligands
HCN
H2N (CH2)6 NH2 NC (CH2)4 CN
Ni cat. with ligands
Hexamethylenediamine Adiponitrile
Monsanto synthesis
H2, eletrochydro– H2
NC CH CH2 + H2C CH CN NC (CH2)4 CN H2N (CH2)6 NH2
dimerization
Acrylonitrile Adiponitrile Hexamethylenediamine
Note that for the conventional method, the byproduct is ammonium sulfate. For
every pound of caprolactam produced, 4.4 pounds of ammonium sulfate is obtained.
The high amount comes from sequential steps: manufacture of hydroxylamine
sulfate, manufacture of the oxime, and Beckman rearrangement.
Toray
NOH·HCl
HCl
+ NOCl
Light
Nitronyl Cyclohexanone
Cyclohexanone
chloride oxime hydrochloride
DuPont
NO2 NOH
H
N
HNO3 Zn+Cr C O
Union Carbide
O
H
O N
CH3COOH NH3
C O C O
Peracetic acid
SNIA Viscosa
CH3 COOH COOH
NH. H2SO4
O2 H2 NOHSO4 C O
+ CO2
160 °C 170 °C in H2SO4
DSM
A buffered solution of NH2 OH, NO3 − + 2H+ → NH3 OH+ + 2H2 O
The buffered NH3 OH+ in H2 O is reacted with cyclohexanone in toluene to pro-
duce the oxime, which undergoes catalytic Beckmann rearrangement to yield final
caprolactam product. The spent aqueous solution is recycled back to previous reac-
tion steps.
ENI
OH
O N
Cyclohexanone Cyclohexanone
oxime
H
N
C O
Caprolactam
HC Catalytic H2 C Co2(CO)8 H2 C
carbonylation Hydroformylation
CH2 CH3 CH2CHO
with double bond shift
CH2COOCH3
H
N –CH3OH H2C NH3, H2
C O –H2O
Ring closure H2C
Catalytic
CH2CH2NH2 reductive
Caprolactam
amination
CN CH2NH2 CH2NH2
CH2
2HCN H2
+
CN CN CH2NH2
CH2
Butadiene Dinitrile Ω-Cyanopentylamine Hexamethylenediamine
(HMDA)
H2O
–NH3
H Nylon 66
N
C O
Caprolactam
Nylon 6
CH3
AlCl3
CH3
+ H3C CH CH2
6.2 Polymer Production from Basic Aromatic Building Blocks 129
Cumene to Phenol
CH3 CH3
O OH
CH3 + O2
CH3
Oxidation
Cumene
[H+] Acid hydrolysis
OH O
+ H3C CH3
Phenol Acetone
The phenol market is globally driven by the demand for bisphenol A (BPA), which
consumed 47% of global phenol production in 2017 (Table 6.4). BPA is driven by
strong demand for polycarbonate products, which outpaced the GDP growth rate
and is expected to continue in the future. Polycarbonate is used in automotive, OEM,
construction, optical media, and appliance industries. It directly competes with glass
and acrylic resins in glazing/sheet and acrylonitrile-butadiene-styrene (ABS) appli-
cations. BPA is also driven by strong demand of epoxy resins, which are used to
make surface coatings, circuit boards, and adhesives. The global use for phenol is
summarized in Table 6.3, which is adapted from 2018 IHS Markit report [3].
Application Examples
Bisphenol 47
Phenol formaldehyde (PF) resin 27
Nylon 15
Alkylphenol 3
Xylenol 2
Other 6
HO
O
[H+] CH3
+ H3C CH3 HO OH
CH3
Bisphenol A
NO2
H2SO4
+ HNO3 + H2O
NO2 NH2
+ 9 Fe + 4 H2O + 3 Fe3O4
HO
NH2
+ NH3 + H2O
NH2
HCHO HCl
+ NH CH2 NH2
p-aminobenzylaniline
Isomerization
O C N CH2 N C O + 4HCl
MDI
(+ poly-MDI)
NO2 NH2
Dinitrotoluene Diaminotoluene
CH3
NCO
+ 4HCl
NCO
Toluene diisocyanate (TDI)
132 6 Uses of Aromatics
Application Percentage-%
+ Polyols
Polyurethane
CH3
H3C CH2 CH2 CH3
(CH2)4 (CH2)5
H3C CH C CH
CH3
CH3 CH3
SO3H SO3H
O
Air
+ 41/2O2 O + 2CO2 + 2H2O
350–400 °C
V2O5 O
Maleic
Glycols anhydride O
Unsaturated O
polyester resins
O
Phthalic anhydride
Cl
HNO3
+ 1/ O2
HCl + 2
–H2O + Trace DCB
Cl
–H2O
2 + CCl3-CHO Cl CH Cl
CCl3
Benzenesulfonic acid
Phenolic resins
Chlorobenzene Phenol
Bisphenol A
Cumene
Benzene Nylon 6,6
80 000 100%
Demand growth CAGR = 6.0%
95%
70 000
90%
60 000
% Capacity utilization
85%
50 000
80
KMTA of pX
40 000 75%
Total capacity
30 000
Total demand 70%
% Utilization
65%
20 000
60%
10 000
55%
0 50%
04
0
00
02
06
08
20
12
14
16
18
10
9
20
20
20
20
20
20
19
19
19
19
19
20
20
20
20
20
CH3
H3C
H3C
CH3
CH3
CH3
The freezing points among xylene isomers are different enough that certain
processes can take advantage for separation, for example, crystallization process.
Other separation processes such as adsorptive separation take advantage of different
adsorptive properties of the isomers (Table 6.6). Once the isomers are separated,
they can be oxidized to give acid or anhydride intermediates to produce polyester
fiber, plastic, plasticizer, and resin materials (Figure 6.3). The worldwide con-
sumption of mixed xylenes in 2017 is shown in Table 6.7 with predominantly high
percentage use of p-xylene, 83% of global xylene demand (IHS Markit [3]).
Table 6.7 World consumption of mixed xylenes in 2017 (thousands of metric tons).
p-Xylene 46 523 83
m-Xylene 598 1
o-Xylene 3 077 5
Ethylbenzene — 0
Solvents and others 6 118 11
136 6 Uses of Aromatics
C8-aromatice
Distillation
To purification
Crystallization
or
or zeolite separation
Isomerization
The major use of p-xylene is to produce PTA, which is then used to produce PET
polymer. The PET polymer is in turn used to make PET solid-state resins and PET
films. The overall consumption for mixed xylenes is strongly tied with the global
polyester industry by the high growth of demand, consumption, and development
for polyester.
The total world producers for mixed xylenes consist of about 170 companies;
among those, the top 10 largest companies account for 45% of the market with
major region based in Asia (Figure 6.3). The top 10 producers are listed in Table 6.8.
Isophthalic acid
CH3 COH
O
m-Xylene
O
CH3
COH
Terephthalic acid
H3C
HOC
p-Xylene O
O
CH2OH COH COOH
O
CH3 CH3 CH3 CH3
O
CH3 H 3C
o-Xylene O O
O O
O
O
O
O Phthalic anhydride
O
V2O5
+ 4.5 O2 O + 2CO2
O
Naphthalene Phthalic anhydride
138 6 Uses of Aromatics
The largest consumption component for phthalic anhydride (PA) is for the pro-
duction of phthalate plasticizers, which consumed about 50% of total use of PA in
2016 (IHS Markit [2]) (Table 6.9). China is the world’s biggest consumer of PA for
plasticizers followed by the Western Europe and the United States.
O O C2H5
C OCH2 CH C4H9
O + 2H C CH CH2 OH + H 2O
9 4
O
O O O
HO OH
O + O +
HC CH HC CH
OOC C O O COOC O O O O O O COOC O
O H C R CH CH2 R CH2 O O
2 2 H2C R CH2 CH2 R CH2 O O H C R CH
2 2
n
O
H2C OH H2C OH
O
O HC OH + HC O R
+
H2C O C R H2C O C R
O
O O
Phthalic anhydride Monoglyceride Diglyceride
O O O O O O
H2C CH CH2 O C O CH2 CH O C O CH2 CH O C O CH2 CH CH2
O O CH2 H2C O O
O O O O O C C O
R R C O C O R R
R R
CH3 CH3
Catalyst
+ 1.5 O2 No further oxidation
–H2O
CH3 COOH
p-Xylene
p-Toluic acid
CH3 CH
Acetic acid
+ HBr
Br( from NH4Br or CoBr2)
Co and Mn Acetates
COOH COOH
Asia/Pacific 80
North America 7
Europe/Africa/Middle East 11
Latin America 2
Polyester fibers 63
Polyester solid-state resin 19
Polyester film 4
PET chip 12
Other 2
COOH CH3
H2, Pd
COOH
COOH
p-Carboxybenzaldehyde p-Toluic acid
p-Xylene Dimethyl
terephthalate (DMT)
Ethylene glycol
Purified terephthalic Terephthalic acid
acid (PTA) (PA)
Poly(ethylene
Ethylene glycol terepthalate) (PET)
Poly(ethylene
terepthalate) (PET)
6.3 Resins and Chemicals Production 141
Poil(butylene terephthalate)
Poly(p-Xylene)
Polyamide resins
Diphenyl
Polybenzimidazoles
isophthalate
Phthalix
o-Xylene Alkyd resins
anhydride
Unsaturated polyesters
Temperature O2 permeation
Material resistance resistance Uses
O O
H O C O CH2 CH2 OH
CH2 OH
O CH2
O
H C
O n
6.4 Summary
Aromatic applications have been contributing to the society by improving the qual-
ity and every aspect of life. Industrial organic chemistry has evolved from the field
of aromatic chemistry giving ways to the development of material science. From
building-block BTX aromatics, many applications have been developed ranging from
pharmaceuticals, dyes, detergents, pesticides, plastics, synthetic fibers, and chemical
resins that can be used in construction and automotive industries.
Polymer science continues to develop with the fundamental chemistry using BTX
aromatics. With liquid-crystallinity characteristic, aromatic-based polymers have
been developed and continually invented to have better performance properties at
lower costs and reduced environmental impacts. The more biodegradable chemicals
or intermediates are the current challenges for development to combat wastes and
environmental footprints. The same is true for pesticides and pharmaceutical
products using aromatic-based molecules. They will be more environmentally
benign, nontoxic and can be used effectively at minimal dosage to protect our
environment and human wellbeing.
Acknowledgments
The author of Chapters 5 and 6 would like to dedicate the work to his wife, Linh
Le, and his son, Bryan Pham, for their love and support. He deeply appreciates
his wife’s tolerance and sacrifice for the family with her everyday works, no mat-
ter how big or small. She is always there with him and his son to take care of them
References 143
for the fun days and the bad days. The author, Trung Pham, is grateful and in debt
to Dr. Santi Kulprathipanja for his leadership and professional advice. With Dr. Kul-
prathipanja’s deep wealth of knowledge and expertise, he has taught a great deal
over the years, and the author feels very fortunate to have him as a colleague and
mentor. Trung thanks many colleagues at UOP, including Stanley Frey, Mary Joe
Wier, Clayton Sadler, James Johnson, Patrick Whitchurch, Patrick Silady, Mark Lap-
inski, Dave Lafyatis, Steve Philoon, Patrick Bullen, Aronson Huebner, Gavin Towler,
David Wegerer, and many others, who have discussed and contributed many ideas
to him over the years. He would also like to thank Mark Goldberg and John Simley
for reviewing these book chapters and providing valuable feedback.
References
Section IV
Industrial Separation
147
Adsorption
or van der Waals forces. In liquid physisorption, the adsorbate can be desorbed by
a desorbent as mentioned earlier. In gas-phase separation, the adsorbate can be
recovered by raising the temperature, reducing the partial pressure of a component,
or in some cases, by using a vapor-phase desorbent. This type of adsorption is
generally used for bulk liquid or gas-phase separation. Liquid adsorption will be
the main subject of this chapter’s discussion. This is because bulk liquid adsorption
has a great impact in refining and petrochemical separations. To demonstrate the
impact of bulk liquid-adsorptive separation, two industrial hydrocarbon separation
technologies, the UOP ParexTM and UOP MX SorbexTM processes, will be discussed
at the end of the chapter.
7.2 Adsorbents
An adsorbent is typically a solid porous material that allows molecules of a gas
or liquid mixture to adhere to its surface, and this process is called adsorption. In
separation processes via adsorption, the role of the adsorbent is to provide a large
surface area for the adsorption of certain molecules. The strength with which these
molecules are bonded to the surface is determined by the nature of the interaction
between the molecule and the surface of the adsorbent. Adsorbents are made from
many types of materials such as zeolites, silica gel, activated alumina, activated clay,
polymeric resin, and activated carbon. General examples of adsorbent applications
are listed in Table 7.1. The selection of adsorbents is dictated by the physical and
chemical characteristics of adsorbents.
As shown in Table 7.1, zeolites are used in industrial bulk-scale separations such as
xylene separation and separation of n-paraffins from i-paraffins. Zeolites are micro-
porous crystalline aluminosilicate minerals with pores of uniform size from 3 to
10 Å, which are uniquely determined by the unit structure of the crystal. These pores
will completely exclude molecules that are larger than their diameter. In contrast,
activated carbon, activated alumina, silica gel, and activated clay do not possess an
ordered crystal structure and consequently the pores are nonuniform. The distri-
bution of the pore diameters within these adsorbent particles may be as narrow as
20–50 Å, or it may range as wide as 20 Å to several thousand Angstrom as is the case
for some activated carbons. In this case, almost all molecular species may enter the
pores. This is the reason why silica gel, activated alumina, activated clay, and acti-
vated carbons are not suitable for industrial bulk separation. Their uses are most
beneficial in the areas of polar species removal from vapor or liquid. The pore size
distribution for zeolite molecular sieve, a typical silica gel, and activated alumina are
illustrated schematically in Figure 7.1.
Zeolites Separate individual PX, MX, or ethyl benzene (EB) from C8-aromatics
Separate n-paraffins from i-paraffins and cyclic
Separate olefins from paraffins
Separate fructose from glucose in corn syrup
Separate individual N2 or O2 from air
Remove acid gas from natural gas
Remove iodide from nuclear off-gas or acetic acid
Remove water from azeotropes, or organic liquids
Purify H2
Silica gel Desiccant in packings and double glazing
Drying of gases, refrigerants, organic solvents
Dew point control of natural gas
Removal of polar impurities from hydrocarbons
Activated alumina Removal of SOx and NOx
Removal of vinyl chloride monomer (VCM) from air
Removal of odors from gases
Clean-up of nuclear off-gases
Purification of helium
Activated clays Removal of organic pigments
Refining of mineral oils
Removal of polychlorinated biphenyls (PCBs)
Removal of dienes/trienes from hydrocarbons
Polymeric resins Water purification
Recover and purify steroids, amino acids
Separate fatty acids from water and toluene
Separate fructose from glucose in corn syrup
Separate aromatics from aliphatic
Recover proteins and enzymes
Removal of colors from syrups
Removal of organics from hydrogen peroxide
Activated carbon Water treatment: portable, industrial, municipal
Sweetener decolorization
Chemical processing
Pharmaceuticals
Hydrocarbon recovery
Industrial off-gas control
150 7 Adsorption
100
3A
90
4A
80
5A
Percentage of pores
70 13X
60 Silica gel
50 Activated alumina
40
30
20
10
0
1 10 100 1000 10000
Angstroms
Raw materials:
(SiO2, Al2O3, Na2O, H2O)
Gel make-up
Heat
Crystallization
Solid/liqiud
separation
parameters have a strong influence on not only the identity of the framework which
is synthesized, but also on the crystal size distribution, the yield of solids, and the
crystallinity of the yielded solids.
Clay
Filter cake
Dryer exhaust
H2O
Dust recycle Gramulator
Weigh
Mixing Vent
Natural gas
Extruder
Natural gas Reject
are the two techniques that are used to manufacture extrudate and bead adsorbents,
respectively.
is like that which is used in palletization. Only minor differences are expected, such
as the moisture content to suit to each system in making a strong adsorbent. The
method of forming the zeolite powder into a bead adsorbent is shown in Figure 7.5.
This includes adding zeolite (filter cake) and clay into a ribbon blender and trans-
ferring the mixture of zeolite and clay to an accretion drum with water addition.
The blended clay zeolite mixture is accreted to form various sizes of bead particles.
The beads are subsequently screened to the desired size particles as product, while
the over- and undersize beads are recycled back to the accretion drum. The bead
adsorbents are then calcined and finally activated to be the adsorbent product.
Liquid adsorption includes two main events: adsorption and desorption. Adsorp-
tion of an adsorbate onto zeolite adsorbent is dictated by the characteristics of the
adsorbate–adsorbent interaction. A zeolitic adsorbent is a crystalline porous solid.
When adsorbent is immersed in a liquid mixture, the porous solid has the capability
to adsorb specific species from the liquid. At equilibrium, the composition of mate-
rial inside the pores differs from that of the liquid surrounding the porous solid. The
amount of adsorbate that can be taken up by the adsorbent (i.e. the adsorbent capac-
ity) is dependent on the nature of the adsorbate–adsorbent interaction, surface area,
and pore volume of the adsorbent.1
In general, a higher adsorbent capacity is seen with higher adsorbent selectivity,
higher adsorbent surface area, and higher pore volume. This phenomenon creates
a higher concentration profile of the more selectively adsorbed component in the
pores of the adsorbent than in the surrounding liquid. The process of adsorbing the
adsorbate into the zeolitic adsorbent is known as the adsorption step.
To desorb the adsorbate from the zeolitic adsorbent, a desorbent is added. A des-
orbent is a suitable liquid that is capable of displacing – or desorbing – the adsorbate
from the selective pores of the adsorbent. The process of recovering or desorbing the
adsorbate from the adsorbent is known as the desorption step. In a chromatographic
liquid separation process, the adsorption and desorption steps must ideally occur
simultaneously. After the desorption step, both the rejected product (product with
lower selectivity, resulting in less adsorption by adsorbent) and the extracted product
(product with higher selectivity, resulting in strong adsorption by adsorbent) contain
desorbent. In general, the desorbent is recovered by fractionation or evaporation and
recycled back into the system.
1 Parts of 7.3 are reproduced from Kulprathipanja [14], Copyright Wiley-VCH Verlag GmbH & Co.
KGaA. Reproduced with permission.
7.3 Bulk Liquid-Adsorptive Separation 155
Ribbon blender
Blower
Dust collector
n
retio
Acc m
d r u
Water
Screen
Hardening
drum
Screen
Over
size Hardening
Beads
drum
RB
Net retention volume of A = RA
RA Net retention volume of B = RB
Selectivity (B/A) = B = RB/RA
Envelope half-width tracer = Wt
Relative concentration
Envelope half-width, B = WB
Rate parameter = WB−Wt = DW
Tracer
A B
Wt WB
Volume (time)
Figure 7.6 Schematic pulse test: two components and paraffin tracer.
(1) Introduce solvent (such as desorbent) into a packed adsorbent column at a fixed
flow rate, pressure, and temperature.
(2) Introduce a feed mixture to be separated until the effluent reaches the feed com-
position.
The feed breakthrough comprises the components to be separated, the tracer,
and the desorbent.
(3) Introduce solvent (such as adsorbent) until no component in the feed mixture is
detected in the effluents.
FRC
UOP rotary valve
1 D 13
12 PRC PRC 24
FR
plot, zeolitic adsorbent selectivity, capacity, mass transfer rate, and desorbent
strength can be calculated.
The UOP rotary valve has been used in more than two hundred Sorbex units across
a variety of applications such as p-xylene (PX) separation, m-xylene (MX) separation,
olefin separation, and n-paraffin separation. The purpose of the rotary valve is to
move the inlet and outlet ports of the net streams (feed, desorbent, raffinate, extract)
around the 24 beds in stepwise fashion, creating a semicontinuous simulated coun-
tercurrent flow of adsorbent relative to the liquid phase. This allows simultaneous
adsorption, purification, and desorption to take place with more than an order of
magnitude efficiency over batch adsorption [19].
exchanged metal cations, SiO2 /Al2 O3 ratio, and water content, can be manipulated
to influence the acidity of zeolites, which affects separation performance. The
degree of separation is characterized by the zeolite selectivity and capacity, as well
as the choice of desorbent and operating conditions such as temperature, feed,
and desorbent flow rates. Therefore, to achieve a meaningful separation using the
equilibrium-selective adsorption mechanism, the following zeolite characteristics
and operating conditions are to be considered:
(1) Zeolite framework structure;
(2) Metal cation exchanged in zeolite;
(3) Zeolite SiO2 /Al2 O3 molar ratio;
(4) Moisture content in zeolite;
(5) Characteristic of desorbent;
(6) Operating temperature.
These variables are described in the following sections.
to alter their acidity and other properties. There is a strong correlation between
the total acidity of a zeolite (the sum of both Brönsted and Lewis acids) and the
ionic radius of the cation as well as the valence charge of the exchanged cation
[21–23]. Exchanged cations with lower ionic radii have higher zeolite acidity. The
correlation between zeolite acidity and ionic radius or exchanged cation valence
is measured by titration with n-butylamine and a Hammett indicator, methylred.
Zeolite acidity increases for monovalent exchanged cations from Cs < K < Na < Li
and increases from Ba < Sr < Ca < Mg and Ba < Pb < La < Ca < Zn < Co for divalent
exchanged cations. For cations of similar ionic radii, divalent cations have higher
zeolite acidity than monovalent cations. For example, Ba-zeolite has a higher acidity
than K-zeolite, and Ca-zeolite has a higher acidity than Na-zeolite and Li-zeolite.
Furthermore, strong acidic Brönsted properties can be observed in hydrated zeolites
due to polarization of adsorbed water in the strong electrostatic field between the
exchanged cations and the AlO4 – anions [23]. For hydrated zeolites, a higher
moisture content results in higher zeolite acidity.
The importance of zeolite acidic strength with zeolite Y for C8-aromatics (xylenes)
systems is illustrated below. In the presence of strong acids, xylene isomers have
varying basicity (Table 7.3), with MX being the most basic and PX the least basic
162 7 Adsorption
Benzene 0.09
Toluene 0.63
p-Xylene 1.0
o-Xylene 1.1
m-Xylene 26
Durene (1,2,4,5-tetramethylbenzene) 140
Isodurene (1,2,3,5-tetramethylbenzene) 16 000
Relative concentration
PX
MX
EB OX
Eluted volume
among the C8-aromatics [24]. Based on the basicity of the xylenes, the acidity of Y
zeolite can be properly adjusted to selectively adsorb MX or PX. As demonstrated in
Figure 7.8, a more acidic zeolite such as NaY will selectively adsorb MX from other
C8-aromatics [25, 26]. In contrast, Figure 7.9 shows that a weaker acidic zeolite such
as KY will selectively adsorb PX from other C8-aromatics [27, 28]. In both systems,
toluene is used as the desorbent. However, acid-based interactions between zeolitic
adsorbents and adsorbates do not always correctly predict the trend of adsorbent
selectivity. This is illustrated by the adsorptive separation of durene from isodurene.
Pulse-test experiments indicated that the adsorbent selectivity for durene/isodurene
increases from KX < NaX < LiX, shown in Table 7.4 [29]. Because isodurene is a
stronger base than durene (Table 7.3), one would expect that the results for adsor-
bent selectivity shown in Table 7.4 would be the reverse order (LiX < NaX < KX).
This counterintuitive trend may be explained by steric hindrances caused by the dif-
ferences in ionic radii of the cations, which could play a role in the adsorbent and
adsorbate interaction.
7.3 Bulk Liquid-Adsorptive Separation 163
OX
Relative concentration
MX
EB
PX
Eluted volume
K 0.59
Na 0.97
Li 3.50
Relative concentration
Isodurene Durene
Eluted volume
Durene
Isodurene
Eluted volume
Figure 7.12 shows separation of fructose and glucose using adsorbents CaY, CaX,
and KX [31]. The combination of the zeolite framework structure and exchanged
cations is instrumental to this intricate mechanism as seen with the separation of
fructose and glucose. Ca2+ forms complexes with fructose and, thus, CaY shows
a high selectivity for fructose over glucose. In contrast, CaX does not exhibit high
selectivity for fructose. However, KX zeolite shows a good selectivity for glucose over
fructose. No simple explanation can be offered for this elaborate system. In the UOP
SarexTM process, Ca-Y zeolite and, later, Ca-exchanged resin were successfully used
as adsorbents that are selective for fructose.
Concentration
G F
adsorbents. (a) Ca-Y adsorbent, (b) Ca-X adsorbent,
and (c) K-X adsorbent.
Concentration
GF
F G
Concentration
Eluted volume
(c)
water content of the adsorbent is adjusted to balance adsorbent selectivity and the
component mass transfer rate.
Table 7.5 Separatio of p-xylene using KY and BaX adsorbents with various desorbents
[38, 39] by pulse tests.
PX net
retention PX/EB PX/MX PX/OX PX stage
Adsorbent Desorbent volume (ml) selectivity selectivity selectivity time (s)
using C8 -aromatic adsorbates with BaX and KY adsorbents. The results in Table 7.5
further emphasize the desorbent characteristic requirement mentioned above.
For instance, phenyldecane is a suitable desorbent for PX separation using BaX
adsorbent. However, phenyldecane is too weak to desorb PX from KY adsorbents.
In contrast, diphenylmethane offers good separation of PX with KY adsorbent, but
not with BaX. With BaX adsorbent, PX is separated from other C8-aromatics using
1,4-diisopropylbenzene, but not with other isomers of di-isopropylbenzene, such as
1,3-di-isopropylbenzene.
Operating Temperature
Another critical variable in liquid-phase adsorptive separation is the operating tem-
perature. Liquid-phase adsorption must be operated at a temperature that optimizes
selectivity and mass transfer rates. Generally, selectivity increases, and transfer rates
decrease at lower temperatures.
Operating Pressure
In general, operating pressure does not affect liquid-phase adsorption. However,
enough pressure must be applied to maintain the system in the liquid phase during
the entire process.
Table 7.6 Energetics and predictions from modeling molecular diffusion in silicalite.
80
3,3,5-TM-C7
70 2,6-DM-C8
Relative concentration
60 1,3,5-TMB
50
2-M-C9
40
30
20
10
0
20 30 40 50 60 70 80 90 100 110 120
Retention volume (CC)
1,3-Diisopropylbenzene
1,4-Diisopropylbenzene
Time
mechanism is obvious when the kinetic diameter of the molecules and molec-
ular sieve pore size opening is compared. n-C5/6 have kinetic diameters of less
than 4.4 Å, which can diffuse freely into the 4.7 Å pores of the CaA molecular
sieve, while non-n-C5/6 have kinetic diameters of 6.2 Å. A commercial example
of shape-selective adsorption is the UOP MolexTM process, which uses a Ca form
A zeolite adsorbent as a molecular sieve to separate C10 –C14 n-paraffins from
non-n-paraffins (aromatics, branched, naphthenes).
Kinetics
Ion exchange involves the formation and breakage of bonds between ions in solu-
tion and exchange sites in a zeolitic adsorbent. The reaction equilibrium of the ion
exchange process depends most significantly on contact time, operating tempera-
ture, and ionic concentration.
170 7 Adsorption
purified para-xylene from the hydrocarbons that were available in gasoline frac-
tions from refineries. Up to that point, fractional crystallization was being used
for purification of para-xylene, with the recovery limited by eutectic composi-
tion to a little over 60%. The UOP Molex process had already been successfully
commercialized for production of high-purity normal paraffins that were used for
production of biodegradable detergents. Unfortunately, the molecular dimensions
of para-xylene and the other feed molecules did not allow molecular sieving to
be used for para-xylene separation. However, the separation stage efficiencies
that could be provided by zeolites were inherently very attractive. When Union
Carbide showed that X- and Y-faujasites could be produced economically at high
purities, UOP intensified its research on competitive adsorption (where essentially
all feed components can access the active sites). Richard Neuzil [46] discovered
that a para-xylene-selective adsorbent could be made by exchanging barium and
potassium onto X- or Y-faujasite, and that it would be possible to use a different
aromatic hydrocarbon to desorb the para-xylene.
This seminal discovery in 1969 led to the UOP Parex process for recovery of
high-purity para-xylene from a mixture of C8 hydrocarbons. The first Parex process
unit was commissioned in 1970 at URBK in Wesseling, Germany, and was designed
to produce approximately 80 000 metric tons per year of 99% para-xylene, with recov-
ery that exceeded 90%. This plant used a mixture of diethylbenzene isomers as the
desorbent. Thus, the extract and raffinate columns had a fairly simple split to make
between the C8 aromatic components of the feed and the C10 aromatic components
of the desorbent. These plants are referred to as “heavy-desorbent Parex” because
the desorbent has a higher boiling point than the feed components. This original
Parex plant was still on stream in 2019. With advances in adsorbent and other PX
technologies from UOP, the capacity of this plant, using the same hardware, is now
more than 250 000 tons/yr. Today, the UOP Parex process accounts for more than
70% of the world’s 50 000 000 tons/yr of para-xylene, as shown in Figure 7.15.
The growth in Parex process PX production between 1974 and 2014 was all with
heavy-desorbent Parex, with steadily higher-capacity plants being built, and with
numerous advances in adsorbent, desorbent, equipment, heat integration, and
control systems being introduced by UOP over this time. Single-train Parex plants
of up to 3 million tons/d of PX are now offered by UOP. Toray tried to use adsorptive
separations for PX with its aromizing process, but that technology suffered from the
use of horizontal adsorbent chambers where bed settling and flow maldistribution
became problematic. In the 1990s, Axens developed the EluxylTM process for PX
production, which uses >100 switching valves to accomplish the SMB that a single
UOP rotary valve can accomplish. Sinopec developed a similar multiple valve
technology in 2010, primarily for deployment within the Sinopec units in China.
Neither of these technologies has garnered the acceptance that has been achieved
by the UOP Parex process.
In the early days of introducing the Parex process to the industry, some prospec-
tive customers voiced concerns about the availability and cost of diethylbenzene as a
desorbent. UOP’s research had identified an alternative adsorbent formulation that
made toluene effective as a desorbent. Four plants of this type were sold. They were
172 7 Adsorption
55
Nameplate capacity, MTA (millions) 50
45
40
35
30
25
20
15
10
5
0
2002
2000
2004
2006
2008
1970
1972
1978
1980
1982
1984
1986
1988
1990
1992
1994
1996
1998
1976
2012
2014
2016
2018
2010
1974
Figure 7.15 The UOP Parex process accounts for more than 70% of the world’s
50 000 000 tons/yr of para-xylene.
called “light -desorbent Parex” because the desorbent has a lower boiling point than
the feed and were licensed. Two of these plants are still operating in this mode, while
the other two were revamped to the heavy-desorbent Parex configuration. For these
light-desorbent Parex plants, the fractionation in the extract and raffinate columns
is more difficult since there is only one carbon number between feed and desorbent,
and the desorbent must be boiled overhead. As a result, the utility consumption
in the extract and raffinate fractionators was much higher for the light-desorbent
compared to the heavy-desorbent Parex plants. For this reason, most customers pre-
ferred the heavy-desorbent Parex configuration, particularly since p-diethyl benzene
is readily available.
In recent years, there has been renewed interest in the energy input per ton of PX.
The regions that show the highest growth rates in demand for PX have also tended
to be the regions where energy costs are relatively high, for example, fuel gas val-
ued above US$10 or even US$15 per million BTU. Many of these locations also have
relatively high costs for electricity. These driving forces have led to significant inno-
vations in PX flowschemes by UOP, with the goal of significantly reducing the fuel
fires and the electricity consumed per ton of PX.
These advances have been achieved through a “holistic” view of the flowscheme
for converting naphtha to PX. To understand the potential for holistic energy inte-
gration (outside the battery limits of the Parex unit), a simplified flowscheme for PX
production is shown in Figure 7.16. Naphtha is fed to a CCR PlatformingTM process
to produce aromatic rings ranging from C6 to C11 in a stream called “reformate.”
The reformate is fractionated into a C7− and a C8+ fraction. The light fraction can
be sent to a SulfolaneTM extractive distillation unit where nonaromatics are rejected
and purified benzene and toluene are recovered. The C8+ fraction is sent to a xylene
rerun column where the C8 aromatics are taken overhead and sent through the UOP
Parex unit, which recovers purified PX. The raffinate from the Parex unit is then sent
7.4 Commercial Bulk Liquid-Adsorptive Process 173
R-254/284/334
ED Raffinate
SulfolaneTM
unit Benzene
CCR
PlatformingTM Reformate TA-32
unit splitter
B T Heavy
aromatics
TatorayTM column
NHT Unit
ADS-47 I-500
A10+
ADS-50 I-600
p-Xylene
Naphtha
ParexTM IsomarTM Light ends
unit unit
Xylene
column Deheptanizer
column
through a UOP IsomarTM unit that re-equilibrates the xylenes, and the isomerate is
then sent to the xylene rerun column for removal of heavy byproducts. These heavy
aromatic byproducts, along with the toluene and A9+ in the reformate, are sent to a
UOP TatorayTM unit that uses transalkylation reactions to create additional xylenes
and benzene. UOP has offered this integrated aromatic technology since the late
1970s with many plants built throughout the world.
Taking this flowscheme as a basis, UOP identified several process intensification
changes that could be brought together to provide more than a 30% reduction in
energy input per ton of PX. This was done without increasing the capital expendi-
ture of the plant. A key element in this solution was a return to light-desorbent Parex,
but with a significantly improved adsorbent, innovative Parex process configuration,
and significantly different fractionation overhead and bottom composition limits
compared to earlier designs. This new technology has been well accepted by the PX
industry, accounting for more than 20 million tons of additional new PX capacity by
2019, a large fraction of the world’s demand for para-xylene [47].
7.4.2 MX SorbexTM
Of the four C8 aromatic isomers, MX is in highest concentration in most feed
streams, amounting to more than 50% of the C8 aromatics. As discussed in Section
7.4.1, MX is sent to the raffinate of the Parex process, and from there it is sent to
the Isomar process where it is converted back to additional PX. Despite being the
dominant C8 aromatic isomer, the end uses for purified MX are comparatively small
relative to those for PX. Whereas PX is primarily used for polyethylene terephthalate
174 7 Adsorption
(PET) fiber, resin, and film in relatively large amounts, the dominant use of MX
is to produce a co-monomer for PET resin called purified isophthalic acid (PIA)
where the co-monomer content is only 2–3% of the resin. Other uses for MX would
be for meta-xylene diamine (MXDA), which can be combined with adipic acid to
produce coatings and gas barriers for containers. The production of high-purity
MX was highly influenced by Mitsubishi and Amoco, which used HF/BF3 adduct
technology to recover and purify MX for many years.
UOP had known how to purify and recover MX for many years using adsorptive
separations [47] and worked steadily to improve the efficiency of this separation. In
the 1990s, the applications for PET resin quickly expanded, with sustained growth
rates above 10%, for carbonated and noncarbonated beverages. The key to this expan-
sion was the abundance or relatively low-cost MX derived from the UOP MX Sorbex
process. Beginning in 1998, UOP has now licensed several of these plants to pro-
ducers in the United States, Asia, and Europe, even to the two companies that had
originally been using HF/BF3 technology.
The MX Sorbex process is very similar to the light-desorbent Parex process in the
sense that the product MX is selectively adsorbed and extracted from the mixed
xylene feed. Toluene is used as the desorbent, with the desorbent recovered in the
overheads of the extract and raffinate columns. Many producers are also in the PX
market and can pull a slipstream of feed from their PX production plants to feed
their (relatively smaller) MX Sorbex plants. In contrast to the 50 million tons/yr of
PX produced in 2018, the worldwide production of high-purity MX is in the range of
1 million tons/yr.
Acknowledgment
The author of the chapter would like to thank James A. Johnson of Honeywell
UOP for writing the “Commercial Bulk Liquid-Adsorptive Process,” which includes
ParexTM and MX SorbexTM in this chapter.
References
1 Li, L., Xue, B., Chen, J. et al. (2005). Direct synthesis of zeolite coatings on
cordierite supports by in situ hydrothermal method. Applied Catalysis A: General
292: 312.
2 Sawad, J.A., Alizadeh-Khiavi, S., Roy, S., and Kuznicki, S.M. (2005). High density
adsorbent structures. W.I.P.O. Patent WO2005032694.
3 Golden, T.C., Golden, C.M.A., and Battavio, P.J. (2005). Multilayered adsorbent
system for gas separations by pressure swing adsorption. US Patent US6893483.
4 Breck, D.W. (1974). Zeolite Molecular Sieves. New York: Wiley.
5 Barrer, R.M. (1982). Hydrothermal Chemistry of Zeolites. London: Academic
Press.
References 175
6 van Bekkum, H., Flanigen, E.M., Jacobs, P.A., and Jansen, J.C. (eds.) (2001).
Introduction to Zeolite Science and Practice, Studies in Surface Science and Catal-
ysis, vol. 137. Amsterdam: Elsevier Science Publishers B.V.
7 Xu, R., Pang, W., Yu, J. et al. (2007). Chemistry of Zeolites and Related Porous
Materials: Synthesis and Structure. John Wiley & Sons (Asia) Pte. Ltd.
8 Thompson, R.W. (1998). Recent advances in the understanding of zeolite synthe-
sis. In: Molecular Sieves Science and Technology: Synthesis, vol. 1 (eds. H.G. Karge
and J. Weitkamp), 1–33. Berlin: Spriger-Verlag.
9 Stiles, A.B. (1987). Catalyst Supports and Supported Catalysts. Boston: Butter-
worths.
10 Stiles, A.B. and Koch, T.A. (1995). Catalyst Manufacture, 2e. New York: Marcel
Dekker.
11 Beck, D.W. (1974). Zeolite Molecular Sieves: Structure, Chemistry, and Use. New
York: Wiley.
12 Shams, K. and Mirmohammadi, S.J. (2007). Preparation of 5A zeolite mono-
lith granular extrudates using kaolin: investigation of the effect of binder on
sieving/adsorption properties using a mixture of linear and branched paraffin
hydrocarbons. Microporous and Mesoporous Materials 106: 268–277.
13 Jasra, R.V., Tyagi, B., Badheka, Y.M. et al. (2003). Effect of clay binder on
sorption and catalytic properties of zeolite pellets. Industrial and Engineering
Chemistry Research 42: 3263–3272.
14 Kulprathipanja, S. (2010). Zeolites in Industrial Separation and Catalysis. New
York: Wiley.
15 Serrano, D.P., Sanz, R., Pizarro, P. et al. (2009). Preparation of extruded catalysts
based on TS-1 zeolite for their application in propylene epoxidation. Catalysis
Today 143: 151–157.
16 Gordina, N.E., Prokof’ev, V.Y., and Il’in, A.P. (2005). Extrusion molding of sor-
bents based on synthesized zeolite. Glass and Ceramics 62 (9–10): 282–286.
17 Ruthven, D.M. (1984). Principles of Adsorption and Adsorption Processes. USA:
Wiley.
18 Neuzil, R.W. (1982). Separation of m-xylene. US Patent 4,326,092.
19 Broughton, D. and Gerhold, C. (1961). Continue separation process employing
fixed bed of sorbent and moving inlets and outlet. US Patent 2,985,589.
20 Namba, S., Kanai, Y., Shoji, H., and Yashima, T. (1984). Separation of p-isomers
from disubstituted benzenes by means of shape-selective adsorption on morden-
ite and ZSM-5 zeolites. Zeolite 4: 77–80.
21 Barthomeuf, D. (1996). Basic zeolites: characterization and uses in adsorption
and catalysis. Catalysis Reviews - Science and Engineering 34: 521–612.
22 Seko, M., Miyake, T., and Inada, K. (1979). Economical p-xylene and ethylben-
zene separated from mixed xylene. Industrial & Engineering Chemistry Product
Research and Development 18 (4): 263–268.
23 Ward, J.W. (1968). The nature of active sites on zeolites. III. The alkali and
lkaline earth ion-exchanged forms. Journal of Catalysis 10: 34–46.
176 7 Adsorption
24 Kilpatrick, M. and Luborsky, I.E. (1953). The base strengths of aromatic hydro-
carbons relative to hydrofl uoric acid in anhydrous hydrofluoric acid as the
solvent. Journal of the American Chemical Society 75: 577.
25 Kulprathipanja, S. (1995). Process for the adsorptive separation of metaxylene
from aromatic hydrocarbons. US Patent 5,382,747.
26 Kulprathipanja, S. (1983). Separation of bi-alkyl-substituted monocyclic aromatic
isomers with pyrolyzed adsorbent. US Patent 4,423,279.
27 Shimura, M., Wakamatsu, S., and Shirato, Y. (1996). Separation of p-xylene by
using zeolitic adsorbents. Japanese Patent 08,217,700.
28 Zinnen, H.A. (1989). Zeolitic p-xylene separation with tetralin heavy desorbent.
US Patent 4,886,930.
29 Kulprathipanja, S., Khune, K.K., and Patton, M.S. (1993). Process for separating
durene from substituted benzene hydrocarbons. US Patent 5,223,589.
30 Robo, J.A. and Gajda, G.J. (1989-1990). Acid function in zeolites: recent process.
Catalysis Reviews - Science and Engineering 31 (4): 385–430.
31 Johnson, J.A. and Kulprathipanja, S. (1989). Proceedings of the International
Conference on Recent Development in Petrochemical and Polymer Technologies
(12–16 December). Bangkok, Thailand: Petroleum and Petrochemical College,
Chulalongkorn University.
32 Kulprathipanja, S. and Neuzil, R.W. (1984). Low temperature process for separat-
ing hydrocarbons. US Patent 4,455,444.
33 Kulprathipanja, S. (2001). Monomethyl paraffin adsorptive separation process. US
Patent 6,222,088 B1.
34 Kulprathipanja, S. (2001). Process for monomethyl acyclic hydrocarbon adsorp-
tive separation. US Patent 6,252,127 B1.
35 Sohn, S.W., Kulprathipanja, S., and Rekoske, J.E. (2003). Monomethyl paraffin
adsorptive separation process. US Patent 6,670,519 B1.
36 Kulprathipanja, S., Rekoske, R.E., Gatter, M.G., and Sohn, S.W. (2003). Proceed-
ings of the Third Pacific Basin Conference on Adsorption Science and Technology,
Kyongju, Korea (25–29 May).
37 Kulprathipanja, S. (1991). Adsorptive separation process for the purification
of heavy normal paraffins with non-normal hydrocarbon pre-pulse stream. US
Patent 4,992,618.
38 Kulprathipanja, S. (2002). Reactive Separation Process. New York, USA: Taylor &
Francis.
39 Bland, W.F. and Davidson, R.L. (1967). Petroleum Processing Handbook. New
York, USA: McGraw-Hill.
40 Pence, D.T. and Macek, W.J. (1970). Silver zeolite: iodide adsorption studies.
The U.S. Atomic Energy Commission, Idaho Operations Office, Under Contract
#AT (10-1)-1230, Nov. 49 Hilton, C.B. (1986) Removal of iodide compounds from
nonaqueous organic media. US Patent 4,615,806.
41 Kulprathipanja, S., Spehlmann, B.C., Willis, R.R. et al. (1999). Method for
treating an organic liquid contaminated with an iodide compound. US Patent
5,962,735.
References 177
42 Kulprathipanja, S., Vora, B.V., and Li, Y. (1999). Method for treating a liq-
uid stream contaminated with an iodide-containing compound using a solid
absorbent comprising a metal phthalocyanine. US Patent 6,007,724.
43 Kulprathipanja, S., Lewis, G.J., and Willis, R.R. (2001). Method for treating
a liquid stream contaminated with an iodide-containing compound using a
cation-exchanged crystalline manganese phosphate. US Patent 6,190,562 B1.
44 Kulprathipanja, S., Sherman, J.D., Napolitano, A., and Markovs, J. (2002).
Method for treating a liquid stream contaminated with an iodide-containing
compound using a cation-exchanged zeolite. US Patent 6,380,428 B1.
45 Kulprathipanja, S., Vora, B.V., and Leet, W.A. (2003). Combination pre-
treatment/adsorption for treating a liquid stream contaminated with an
iodine-containing compound. US Patent 6,506,935 B1.
46 Neuzil, R.A. (1971). Aromatic hydrocarbon separation by adsorption. US Patent
3558730 A.
47 Frank (Xin X.) Zhu, James A. Johnson, David W. Ablin, and Gregory A. Ernst
(2020). Efficient Petrochemical Processes – Technology, Design and Operation.
Hoboken NJ: Wiley.
179
Distillation
rest flows back into the column and moves downwards as liquid in the opposite
direction of the rising vapor. The upper portion of Blumenthal’s column incor-
porated bubble-cap trays, while the lower part included conical metal caps for
contacting vapor and liquid. These advances in distillation not only helped supply
the refined sugar Europe needed on the less volatile side of this process but also
launched, on the distillate side, the French rum industry.
The coal industry also drove innovations in distillation technology through the
nineteenth century. In 1823, Friedlieb Runge discovered phenol and aniline in coal
tar and, through further distillation, enabled the use of these tar byproducts in col-
orants and pharmaceutical product development. In addition, ammonia, benzene,
and other aromatic compounds beyond phenol were isolated from the coke gas and
were separated by washing, absorption, and distillation. Column and tray designs
continued to improve throughout the nineteenth century as a wider range of dis-
tilled products were produced. The residue from these processes found use in street
paving by the mid-1800s.
The next great leap in distillation technology arrived with the birth of the
petroleum industry. While the first generation of crude oil distillation was a simple
batch process to produce lamp oil, the outbreak of World War I required larger
volumes of oil and fuel to be isolated. With no chemical conversion, only 20–25%
of crude oil could be distilled into fuel and oil products. In the 1910s, thermal
cracking was utilized to generate lighter oils from the longer-chain hydrocarbons
in the crude. In 1916, steam-assisted distillation of petroleum was accomplished,
and the 1920s ushered in a switch in gasoline production from batch to continuous
distillation. Technologies such as thermal cracking via the Dubbs process [3]
allowed much higher yields of gasoline and other smaller-chain hydrocarbons to be
derived from crude oil.
These distillation process advances were paired with a greater understanding of
the theory and design of distillation. The measurement and prediction of mixture
properties, especially vapor–liquid equilibrium (VLE), are crucial to the accurate
prediction of separation performance.
For example, Rayleigh pioneered the use of Henry and Raoult’s laws to compare
test results to data calculations for simple batch distillations. Hausbrand [4] then
advanced this work using graphical representations of the VLE properties for non-
aqueous and aqueous mixtures. In this manner, the beginnings of VLE theory were
established, based on relative volatility (RV) differences in the materials being dis-
tilled. Finally, McCabe and Thiel developed their diagram to simplify the correlation
of distillation column design with VLE properties and mass balance data [5].
For more than 5 000 years, distillation has served as the primary method for the
separation of complex mixtures into refined, purified products. Even today, distilla-
tion is the most widely practiced separation technology and is responsible for approx-
imately 50% of capital and operating costs in industrial chemical processing. Dis-
tillation accounts for 50% of the annual process energy needs of the chemical and
refining industries [6].
Although the refineries of today – as well as those in the future – are far more
complex, with higher yields, higher value creation, lower regrettable emissions, and
8.2 Principles and Systems 181
In this way, K a represents the tendency of component a to volatilize into the vapor
phase and is a function of temperature, pressure, and, to a lesser degree, composi-
tion. If K a equals 1, there is an equal amount of component a in the vapor and liquid
phases, while K a > 1 means more a in the vapor phase and K a < 1 means more a in
the liquid phase [7].
To separate components via distillation, we must understand their relative volatil-
ity. The relative volatility of components a and b is defined as K a /K b . This ratio of
volatility is a measure of how easily two components can be separated by distillation.
If the ratio is high, component a is much more volatile than component b, and the
distillation is straightforward. As the ratio (K a /K b ) approaches a value of one, it is
more difficult to separate each of the components. In the extreme case of the value
of K a /K b being exactly one, separation by distillation is not possible other than by
azeotropic or reactive distillation methods. It is not possible to have a relative volatil-
ity ratio less than one as we always define the ratio as the K value of more volatile
component over the K value of the less volatile component.
182 8 Distillation
Feed, xF
Liquid, La, xa
The McCabe–Thiele design method provides a means to model the vapor and
liquid composition (mole fraction) of a component of a binary mix within a distilla-
tion column. It affords a graphical visualization of distillation principles while also
providing a solution to the material balance and equilibrium relationships. This rel-
atively simple approach allows the designer to build on the fundamentals of VLE
theory and clarify design basics such as the number of trays required to achieve a
given binary component separation as well as where to locate the feed within the
tray structure.
Figure 8.1 depicts a binary distillation of a feed of composition xF into a vapor
phase where the mole fraction of the more volatile component, a, is designed as ya
and the mole fraction of a in the liquid phase is xa .
Figure 8.2 shows the equilibrium of a single stage of the distillation of this feed
(labeled stage 1). Because the system is in equilibrium, there are two conditions that
are required to satisfy VLE:
(1) The amount of component a leaving the stage (La xa + V a ya ) is equal to the
amount of vapor and liquid entering the stage from the trays positioned before
(V a−1 ya−1 ) and after stage 1 (La+1 xa+1 ), respectively.
(2) The amount of component a in the vapor phase at equilibrium is equal to the K
factor times the mole fraction in the liquid phase, ya = K a (xa ).
Va–1 La
ya–1 xa
8.2 Principles and Systems 183
0.4 5
T5
0.2
ya = 0.9
xa = 0.1 xF = 0.5
0
0 0.2 0.4 0.6 0.8 1.0
Liquid mole fraction
(1) Label the axis as the mole fraction of the lighter component (a) in the vapor
phase (y-axis) versus the mole fraction of the same lighter component (a) in the
liquid phase (x-axis).
(2) Plot Line 1 with slope 1 and intercept of 0 across the diagram.
(3) Plot Curve 2: the VLE curve. This curve is derived from primary thermodynamic
reference values. Recall that according to VLE, ya = K a (xa ).
(4) Determine the feed composition: In this example, the feed contains 50% compo-
nent a, that is, xF = 0.5.
(5) Plot Line 3: the quality line or q-line, which is shown in Figure 8.3 as a mixture
with equal mole of component a as vapor and liquid in the feed. The line, there-
fore, has a slope of −1. M = q/(q − 1), where q = feed quality (liquid mol fraction
of the feed).
(6) Plot Line 4: the rectifying line. For a distillate composition of 90% component a,
we start this line at 0.9 on the x-axis and intersect Line 1. The slope of this line is
L/(L + D), where L = the molar flow rate of reflux and D is the molar flow rate of
the distillate product. The line is extended until it intersects Line 3 (the q-line).
(7) Plot Line 5 (stripping line). For a bottoms composition of 10% component a,
we start this line at 0.1 on the x-axis and intersect the point where Lines 1 and
3 meet.
(8) Now, we can plot the tray structure for this distillation by starting at the xa = 0.1
point and creating a stairstep structure, which stops when it intersects either the
stripping or rectifying lines.
More
volatile
T1 product, ya
T2 Rectifying
section
Feed, xF
T3
T4
Stripping
T5 section
Reboiler
Less
volatile
product, xa
within the tray structure. The intersection point of the rectifying line and stripping
line falls with the Tray 3 step; therefore, Tray 3 is the best location for the feed in this
distillation setup. The resulting column and feed setup is shown in Figure 8.4.
Boiling point is not a static value. Rather, the boiling point of a mixture will vary
as temperature, pressure, and composition change across the distillation column.
As required by VLE theory, the vapor and liquid within each stage reach equilibrium
and no further separation is possible; each stage reaches a unique equilibrium con-
centration of the mixture components according to their relative volatility. Further
purification of the distillate (more volatile components) or bottoms (less volatile
components) is possible through the introduction of either more distillation columns
or other separation methods such as absorption or filtration.
Mixtures with a relative volatility (RV) value greater than 1.5 are described as a
simple distillation and require no special measures to accomplish the desired sep-
aration. By contrast, mixtures with relative volatility (RV) values less than 1.2 are
described as extractive distillation; these separations can be eased by the addition
of heavier components. This method will be described in more detail later in the
chapter [9].
8.2.3.3 Solubility
Solubility is a function of temperature and pressure, so the relative solubility of com-
ponents in the feed as well as throughout the distillation column can make it more
or less difficult to separate those components. Differences in solubility can be used,
for example, to make possible the separation of materials that have the same rela-
tive volatility through the addition of a third component with a higher solubility for
one of the components; this method is referred to as azeotropic distillation and is
described in more detail later in this chapter.
this example is volatilized under the same pressure and temperature, it will yield the
same concentration of components a and b. In this case, it is not possible to further
separate these components as they have formed an azeotropic mixture that has hit
a constant, maximum concentration limit. One simple example of this maximum
concentration behavior is isopropyl alcohol in water; the concentration of isopropyl
alcohol cannot rise above 91% as the boiling points for the two components are equal
at this maximum concentration limit.
● Tray hydraulics: Tray hydraulics and component rheology dictate the flow behav-
ior of a liquid through a tray distillation system. These material properties, along
with the tray design properties, will have a significant impact on the ability of the
distillation apparatus to create the surface area needed to drive separations.
● Feed component stability: The feed boiling point needs to be compared to the indi-
vidual component’s decomposition temperature. Additionally, there is a need to
understand any possible reactions that may occur between components within the
distillation column as relative volatilities for these components can vary greatly
from the feed mixture components.
selected when there are drivers other than cost to consider for a separation process.
Some of the advantages of batch distillation include:
(1) Simplicity: Multiple components can be separated using only one feed and only
one distillation column.
(2) Flexibility: One batch distillation column can be used to handle a wide range of
feed materials with varying quality and composition.
(3) Capacity: Batch distillation is well suited for small volume separations.
(4) Feed difficulty: If a feed has high viscosity or high solids levels, it is easier to run
a batch distillation than use continuous processing methods.
(5) Quality control: Batch distillations can be quality-verified easily based on both
feed and product specifications.
Continuous Distillation
Continuous distillation processes complement batch processing in that they are
lower cost and are tuned to one specific type of feed. The emphasis for continuous
distillations is on stability and the production of a high volume of product stream(s)
at consistent quality.
When constructing a continuous distillation strategy, there are several column
options to consider:
(1) Binary column: Feed contains two components.
(2) Multicomponent column: Feed contains more than two components
(e.g. crude oil).
(3) Multiproduct column: Column has more than two product streams located at
different points along the height of the column.
(4) Extractive distillation: Where feed appears in the bottom product stream.
(5) Azeotropic distillation: Where feed appears at the top product stream.
Horizontal flow across the trays Vertical flow through the packing
(counter-current)
Large diameter columns and with high Small-diameter (D < 0.6 m) columns with
liquid flow rates (>30 m3 /(m2 h)) low liquid flow-rate applications
(<50 m3 /(m2 h))
Metal Metal, ceramic, and plastic
High liquid residence time Low liquid residence time (i.e. for
heat-sensitive materials)
High-pressure drop (7 mbar per Low-pressure drop (0.1–0.5 mbar per
equilibrium stage) equilibrium stage) as required for vacuum
distillation
Easy to clean Not typically cleaned/reused
Production 1
Reflux distillate
Rectifying
section
Feed
Stripping
section
Reboiler
Product 2
bottoms
8.3 Distillation Columns and Trays 189
a continuous distillation tower containing nine trays with the binary feed mixture
brought into contact with Tray 6. Starting at the top of this column, we see that part
of the condensate is returned to the column as reflux material to provide liquid flow
downward through the column. Likewise, a portion of the liquid bottoms fraction is
returned through the reboiler to the column and provides the upward flow of vapor
from the bottom of the column. The stripping section (below the feed inlet) primarily
removes the more volatile components from the liquid, while the rectifying section
(above the feed inlet) includes an enriched level of volatile components. Once the
targeted level of separation is achieved, this binary, continuous distillation results in
two products, the distillate (more volatile components) and the bottoms (less volatile
component).
As described earlier, binary systems in which the relative volatility is greater than
1.5 fall in the category of simple distillations and, in some cases, the rectifying section
of the column in Figure 8.5 can be removed to yield a simpler stripping column.
Product 1
Reflux
Feed 1
Product 2
Feed 2 Product 3
Reboiler Product 4
Product 5
190 8 Distillation
Figure 8.8 Common distillation tray types. (a) Sieve, (b) Valve, and (c) Bubble-cap.
(2) Operating capacity: The diameter of the column required to provide a give flow
rate is largely invariant across the three types of trays. Sieve trays have a slightly
higher capacity, then valve, and finally bubble-cap.
(3) Cost: Bubble-cap trays are more complex and, therefore, more expensive than
sieve or valve trays.
(4) Pressure drop: As might be imagined, sieve trays have the lowest pressure drop
followed by valve and then bubble-cap trays. Given the liquid low properties of
the feeds involved, vacuum columns need to account more for pressure drop
than atmospheric distillation columns.
Shown in Figure 8.8 are the three types of distillation trays we are most interested
in (sieve, valve, and bubble-cap). Each of the three types is described below:
(a) Sieve tray: This tray is simply a flat plate with 5–6-mm-diameter holes (or slots)
punched in it. The liquid flowing across the holes and tray is retained by vapor
upflow from below. Due to its simplicity, sieve trays are the least expensive, have
the lowest pressure drop, and are most resilient to solids and corrosion (espe-
cially when large holes are used). For efficient operation, the velocity of vapor
through the sieve tray holes must be sufficiently high to prevent liquid on the
tray deck from passing through the perforations to the tray below (weeping).
Conversely, if the velocity of the vapor through the sieve tray perforations is too
high, liquid from the tray below will transfer to tray above (entrainment) and
will cause lower column efficiency. Because of these limitations, the turndown
performance for sieve trays is relatively poor at 2 : 1.
(b) Valve tray: A valve tray is similar in construction to the simpler sieve tray, but
includes a moving valve (e.g. disk), which closes when vapor-flow rates are low
to avoid weeping of the liquid to the stage below. For this reason, valve trays are
characterized by higher turndown ratings (5 : 1 range) due to improved perfor-
mance at low vapor-flow rates but are more costly than sieve trays (about 1.2
times more expensive).
(c) Bubble-cap tray: A bubble-cap tray has riser or chimney fitted over each opening,
and a cap that covers the short riser pipe. The cap is carefully placed to maintain
a space between riser and cap, which allows the passage of vapor. This design
is the oldest type of cross-flow tray and ensures that liquid is maintained on the
tray at all vapor flow rates. While bubble-cap trays bring obvious advantages in
turndown performance (10 : 1 and higher), they suffer from high pressure drops,
high cost (2× sieve trays), and poor corrosion resistance.
A comparison of these three tray types is included in Table 8.2 below.
192 8 Distillation
While distillation itself dates back thousands of years, packed columns only came
into wide use in the 1930s with the introduction of random, but regular, packing
materials such as Raschig rings. Shown in Figure 8.9 is a packed column such as
that used in distillation or absorption; liquid and vapor flow counter-current within
this simple yet effective design.
Packed columns are commonly used for distillation, absorption, stripping, and
heat-exchange applications. The design of these systems is a function of their capac-
ity, pressure drop, and resistance requirements (both mechanical and chemical). The
packing materials loaded with the column establish the capacity and separation effi-
ciency of these distillation systems and have been steadily increasing in performance
over the past several decades to include not only high-performance random packings
but also structured packing materials [10].
Reflux
Packing
Support
Packed bed
Feed
Packing
Support
Liquid Vapor
out in
8.4 Distillation Packing Materials 193
D E
intermingling of the ascending gases and vapors and the descending liquid is thus
affected” [11].
Raschig rings were just the start of a new, improved suite of distillation options.
Shown in Figure 8.11 are a range of packing materials used for packed column dis-
tillations and absorption processes today. Pall rings (D) represent a variant of the
original Raschig ring cylinder (A) design with strips of material folder into the inte-
rior of the cylinder; this change provides more even liquid distribution within the
packing phase.
Subsequent generations of packing materials included Berl and Intalox® saddles
(B and C), which are typically made of ceramics and exhibit superior wetting charac-
teristics paired with high corrosion resistance. The saddle packings were developed
to provide improved liquid distribution versus Raschig rings, and the Intalox sad-
dles were easier to manufacture than the original Berl design. While the saddles
themselves are more expensive than Raschig rings, the overall efficiency of Berl and
Intalox saddles as well as Pall rings has led to them being favored for most applica-
tions today.
Developed in the late 1970s, the Intalox IMTP random packing (E) repre-
sents a third-generation packing design, which combines the advantages of the
saddle-shaped packing with that of modern high-performance ring-type packings.
This shape delivers a lower-pressure drop at the same vapor and liquid loads
compared to previous-generation packings as it is one of the most popular random
packings in use today.
Metal
mesh
Vapor
openings
Separator
sheet Vapor
openings
(separator)
of triangular cells of regular shape and angle. The Stedman column design provided
lower holdup, low pressure drop, shorter column height, and low weight. These
same advantages drive the selection of structured packing today.
A second generation of structured packing designs and materials was introduced
in the 1970s and consisted of corrugated sheet packing, which further raised col-
umn capacity while lowering costs and increasing plugging resistance of the system.
These second-generation options are now widely adopted as attractive options for
column revamps and have in common design elements such as open honeycombed
structures with 45∘ inclined packing angles in combination with perforations to
maximize capacity and minimize pressure drop (Figure 8.12). The purpose of the
angles and design specifications in a structured packing is to force liquids to take
tortuous paths through the column, thereby creating a large surface area for contact
between different phases and increasing the separation efficiency.
Some common materials used in the design of structured packing are described
below:
Sheet metal: Can tolerate modest levels of fouling and handle a wide range of vapor
and liquid rates. Process applications range from low-pressure drop vacuum and
atmospheric distillation services to medium high-pressure absorbers.
Gauze: Preferred packing for deep vacuum and low liquid rate applications as it has
the lowest pressure drop per theoretical stage, which helps in the purification of
temperature sensitive materials such as those found in specialty chemicals and
pharmaceuticals.
Grid: Both corrugated metal and stamped blades are grid-type packing and both
bring with them a high level of tolerance to fouling as they incorporate open vol-
ume to facilitate flow even in the presence of fouling. Corrugated grid packing has
a smooth surface to minimize fouling and is constructed of much thicker material
than the normal sheet metal packing.
Crude
oil Condenser
LPG (C3–C4)
Reflux
Desalter
Pentane / Hexane
(C5–C6)
Naphtha (C6–C11)
Heater
Kerosene Light
(C11–C14) vacuum
gas oil
Diesel
(C11–C20)
Heavy
vacuum
Atoms gas oil
gas oil
(C14–C20)
Reduced
crude Wax
Vacuum
Product 5 Heater cilumn
bottoms
Before distillation can proceed, there is typically a need to clean up the crude oil
as crude can include many problematic materials such as sand, inorganic salts, poly-
meric impurities, and more. In most crudes, NaCl, MgCl2 , and CaSO4 form droplets
dispersed throughout the crude. If not removed, these impurities can cause issues
downstream in the refining process such as abrasion from sand, as well as acidic
gases (e.g. HCl) upon heating of the impure crude.
The most common method to desalt crude oil is electrostatic precipitation
wherein the crude is first heated to decrease its viscosity. Fresh water is introduced
to the heated crude and forms small droplets within the continuous oil phase.
This water-in-oil dispersion is then introduced to a pressurized desalter vessel,
which includes a high-voltage electrical field that accelerates separation of the salt
containing water droplets from the desalted oil. As the separation continues, the
water droplets sink within the oil phase and are removed continuously from the
bottom of the desalter, along with any collected sediment, for disposal. Desalted
oil flows continuously from the top of the pressurized vessel and can then be
introduced to the distillation units. Desalting can typically achieve 99% removal
rates for common salts such as NaCl, MgCl2 , and CaSO4 .
Once the desalting process has removed inorganic salts, the resulting crude is
heated to approximately 360–400 ∘ C in tube furnaces and fed into the distillation por-
tion of the refinery. Most refineries employ two types of distillation columns in series:
8.5 Commercial Examples 197
value of the bottoms products from atmospheric distillation in end markets such as
olefins and aromatic petrochemicals as well as fuels.
Thermally Mechanically
integrate integrate
A A
A B
A/B
Dividing
wall
ABC ABC ABC PF
C1 C2 PF MC B B
B/C C B/C
C C
have been known for many decades. For example, dividing wall columns (DWC)
appeared in patent literature as early as 1946 [15]. Deployment of DWC technology
took place starting in the mid-1980s as control of these systems advanced greatly
with the advance of computer-aided process management tools. The roughly 30%
savings in capital investment and operating expenses have led to more than 200
dividing wall column installations globally in 2020.
Shown in Figure 8.14 is a representation of the sequential separation of a
three-component feed in two columns alongside the integrated dividing wall
column design. In the pre-fractionator section of the dividing wall column, a
rough separation is carried out between components A and C. Component A is
concentrated at the top of the prefractionator and C at the bottom with component
B distributed between the top and the bottom. The pre-fractionator overhead then
feeds the upper half of the main fractionator for the binary separation of A and B,
while the bottom of the pre-fractionator feeds the lower half of the main column for
the separation of B and C.
The main fractionator liquid from the top section refluxes the pre-fractionator top
and the main fractionator on the other side of the dividing wall, while the vapor from
the bottom section similarly strips the bottom of the pre-fractionator and the main
fractionator. The middle component B is withdrawn at the stage where its concen-
tration is at a maximum, avoiding re-mixing losses, which can lead to energy savings
of up to 25–30% or more.
There are now many relatively well-known potential applications for dividing wall
columns, including straight-run naphtha fractionation upstream or downstream of
reformers, FCC naphtha fractionation, C4 isomer separation, and benzene, toluene,
and xylene (BTX) separation. A dividing wall column also is used for UOP’s Pacol
enhancement process (PEP). In this process, A is pentane, B is benzene, C consists
of C7 + olefins, and D comprises C7 + aromatics. To prevent mixing of C and D, a
novel trap tray is applied and excessive B is added [16]. The next logical step in
DWC performance is to include multiple dividing walls within the DWC column,
200 8 Distillation
and designs have been developed to produce six high-purity products from a single
column [17].
References
9 Gerster, J.A., Mertes, T.S., and Colburn, A.P. (1947). Ternary systems
n-butane–1-butane–furfural and isobutane–1-butane–furfural. Industrial & Engi-
neering Chemistry 39 (6): 797–804.
10 Chilton, T.H. and Colburn, A.P. (1935). Distillation and absorption in packed
columns. Industrial & Engineering Chemistry 27 (3): 255–226.
11 Raschig, F. (1915). Absorption and reaction tower for acids. US Patent 1,141,266,
26 filed March 1914 and issued 1 June 1915.
12 Stedman, D.F. (1935). Packing for fractionating columns and the like. US Patent
2,047,444, filed 14 January 1935 and issued 14 July 1936.
13 Alfke, G., Irion, W.W., and Neuwirth, O.S. (2007). Oil refining. In: Ullmann’s
Encyclopedia of Industrial Chemistry (ed. James G. Speight), 7e, 208–260. Wein-
heim: Wiley-VCH.
14 Speight, J.G. (2007). The Chemistry and Technology of Petroleum, 4e. New York:
CRC Press.
15 Wright, R.O. (1946). Fractionation apparatus. US Patent 2,471,134, filed 17 July
1946 and issued 24 May 1949.
16 Yildirim, O., Kiss, A.A., and Kenig, E.Y. (2011). Dividing wall columns in chemi-
cal process industry. Separation and Purification Technology 80: 403–417.
17 Piszczek, R., Hergenrother, M., and Wang, Z. (2020, 2020). Introduction to
dual-dividing-wall columns. Chemical Engineering Progress 4: 22–25.
18 Tower, G.P. and Frey, S. (2002). Reactive distillation. In: Separation Processes
(ed. S. Kulprathipanja), 18–50. New York: Taylor & Francis.
19 Sakuth, M., Reusch, D., and Janowsky, R. (2007). Reactive distillation. In:
Ullmann’s Encyclopedia of Industrial Chemistry (ed. James G. Speight), 7e,
264–275. Weinheim: Wiley-VCH.
203
Membranes
Membrane technology has been an important enabler for effective and efficient
gas, vapor, and liquid separations and a wide range of other potential applications
such as for catalytic membrane reactors, membrane distillation, and energy storage.
This chapter covers perspectives of membranes from a combination of both
academic research and industrial development.
petroleum producers and refiners, chemical companies, and industrial gas suppliers.
Several applications of gas separation membranes have achieved commercial
success, such as nitrogen enrichment from air, carbon dioxide removal from natural
gas and from enhanced oil recovery, separation and recovery of organic vapors, and
in hydrogen removal from nitrogen, methane, and argon in ammonia purge gas
streams. However, membrane technologies with limitations for organic vapor/vapor
or liquid feed separations, such as for olefin/paraffin, aromatic/nonaromatic, and
aromatic/aromatic separations, have not been commercially successful so far.
Polymeric membranes often suffer from thermal stability and plasticization or
swelling of the organic polymer matrix by the sorbed penetrating molecules such
as high concentration of carbon dioxide or olefins, which results in the decrease
of membrane selectivity and therefore lower product recovery and/or higher com-
pression cost due to higher recycle rates. Inorganic membranes such as microp-
orous zeolitic membranes can overcome some of the challenges facing polymeric
membranes. For example, zeolite membranes have the potential for separations with
high efficiency and high productivity under conditions where polymeric membranes
cannot survive by taking advantages of their superior thermal and chemical stabil-
ity, good erosion resistance, and high plasticization resistance to condensable gases
or vapors. The main challenges facing inorganic membranes for large-scale indus-
trial applications are their high cost, reproducibility, difficulty of making very thin
(<100 nm), grain boundary-free, pinhole-free, and crack-free selective layer.
The significant growth of membrane technologies in refinery, petrochemical,
and natural gas markets requires breakthrough development of new membranes,
modules, and processes that overcome the challenges facing the current membrane
technologies. Innovation in membrane materials, equipment, and process technol-
ogy integration will play a key role to enable new large-scale, low-energy-intensity
membrane-based separations.
The solubility coefficient equals the ratio of sorption uptake normalized by some
measure of uptake potential, such as partial pressure. The diffusivity coefficient is a
measure of the mobility of component A or B passing through the voids between
the polymer chains in the polymeric membrane. Therefore, the selectivity of the
polymeric membrane for component A over component B (𝛼 A/B ) is the product of
the solubility selectivity of component A over component B (abbreviated as SA /SB )
and diffusivity selectivity of component A over component B (abbreviated as DA /DB )
(𝛼 A/B = SA /SB × DA /DB ). A polymeric membrane can have a high PA because of high
SA , high DA , or both, and the membrane can have a high 𝛼 A/B because of high SA /SB ,
high SA /SB , or both. Normally the solubility coefficient increases and the diffusiv-
ity coefficient decreases with an increase in the molecular size of the permeating
component for a glassy polymeric membrane and vice versa for a rubbery polymeric
membrane. Both permeability and selectivity are the intrinsic properties of the poly-
meric membrane materials and ideally are constant with the change of the thickness
of the membrane selective layer, feed pressure, flow rate, and other separation pro-
cess conditions although they are temperature dependent.
Membrane permeance, measured in gas permeation units (GPUs,
1 GPU = 10−6 cm3 (STP)/cm2 s (cm Hg)), is the pressure-normalized flux and
determined by the intrinsic permeability of the polymer material and the selective
skin layer thickness of the membrane. Therefore, a commercially viable polymeric
membrane based on a solution-diffusion separation mechanism needs to have a
high intrinsic permeability and a thin selective layer to achieve a sufficiently high
permeance at the preferred separation conditions. Otherwise, extraordinarily large
membrane surface areas are required to achieve high treating capacity. For mem-
brane separation applications, both high permeance and selectivity are desirable
because higher permeance decreases the size of the membrane area required to
treat a given volume of feed mixture, thereby decreasing the capital cost of the
membrane system, and higher selectivity results in a higher-purity product.
Polymeric materials have been studied extensively for the development of mem-
branes for separation applications, particularly for solution-diffusion-mechanism
based separations. Many research articles and patents have reported the synthesis of
polymeric membrane materials, such as polyimides (PIs), polyethersulfones (PESs),
polyethers, polyamides, polybenzoxazoles, polybenzimidazoles (PBIs), and poly-
mers with intrinsic microporosities. Traditional polymeric membrane materials,
however, have shown a well-known tradeoff between permeability and selectivity
(or so-called polymer upper-bound limit). By comparing the permeation data of over
three hundred different polymers, Robeson reported in 1991 and 2008 that the selec-
tivity and permeability of polymeric gas separation membrane materials were insep-
arably linked to one another, in a relation where selectivity increases as permeability
decreases and vice versa [9, 10]. Substantial research effort has been directed to
design novel polymer structures or modify existing polymer structures to overcome
the permeability and selectivity limits imposed by the polymer upper-bound limit
in academia and industrial research labs. Fabrication of defect-free high selectivity
membranes with a thin selective skin layer, development of advanced membrane
9.2 Types and Preparation of Membranes 207
(a) (b)
(c) (d)
(e)
(a) (b)
(c) (d)
Figure 9.2 Cross-sectional HRSEM images of: (a) asymmetric integrally skinned flat-sheet
membrane; (b) TFC flat-sheet membrane; (c) asymmetric integrally skinned hollow-fiber
membrane; and (d) TFC hollow-fiber membrane.
be used as the support membranes for the fabrication of RO, NF, and gas separation
membranes. This chapter will not cover the preparation of this type of asymmetric
integrally skinned porous polymeric membranes in detail.
Asymmetric polymeric membranes for gas separation, RO, NF, and pervaporation
applications can have either an asymmetric integrally skinned membrane structure
or TFC membrane structure. Such membranes normally consist of a thin, dense,
nonporous, selective skin layer of <100 nm and a less-dense void-containing
(or porous), nonselective support layer, with pore sizes ranging from large in the
support region to very small proximate to the skin layer. The thin, dense, nonporous,
selective skin layer provides the membrane resistance and selectivity and the thick,
porous support layer provides the membrane with high mechanical strength.
Design and selection of proper polymeric materials with desired properties is
a critical initial step for the development of asymmetric polymeric membranes.
In addition to the consideration of high permeability and selectivity, the polymer
cost, purity, molecular weight, commercial availability or manufacturing possibility,
solubility of the polymer in organic solvents for the preparation of polymer solu-
tions, film-forming property and membrane processability, mechanical property,
post-treatment possibility, chemical and thermal stabilities, and hydrophilic-
ity/hydrophobicity are all the factors that need to be considered to select the best
9.2 Types and Preparation of Membranes 209
is formed after the organic solvents and nonsolvents diffuse out of the interior of
the layer of polymer solution and the nonsolvent liquid in the coagulation medium
such as water diffuses into the interior of the layer of polymer solution; (v) washing
the nascent asymmetric integrally skinned membrane comprising a thin, dense,
nonporous, selective skin layer and a highly porous support layer in a nonsolvent
medium such as a water bath to remove the leftover organic solvents, nonsol-
vents, and additives from the nascent asymmetric integrally skinned membrane;
(vi) annealing the nascent asymmetric integrally skinned membrane comprising a
thin, dense, nonporous, selective skin layer and a highly porous support layer in an
annealing medium with a controlled temperature such as a hot water-annealing tank
controlled at 55 ∘ C. The trace amount of leftover organic solvents, nonsolvents, and
additives from the coagulating medium will be further removed during this anneal-
ing step. In addition, both the thin, dense, nonporous, selective skin layer and the
highly porous support layer will be further stabilized with certain degree of shrink-
age. The stabilization of the membrane structure is critical to provide the membrane
long-term performance stability for high-pressure applications. This annealing step
is unnecessary for fabricating certain asymmetric integrally skinned membranes;
and finally, (vii) drying the wet asymmetric integrally skinned membrane to remove
the coagulation and annealing fluids from the membrane without causing changes
to the substructure of the membrane at certain temperature from about 50 to about
100 ∘ C. Some membranes need to be kept wet without drying before use and this
drying step can be avoided and replaced with an additional conditioning step by
placing the wet membrane in a conditioning medium that may contain water and
conditioning agent such as glycerol or polyethylene glycerol. The conditioning
agent diffuses into the pores of the membrane to prevent the collapse of the pores
during storage. In some cases, some defects on the thin, dense, nonporous, selective
skin layer of the membrane caused by gas bubbles, dust, or substrate imperfections
cannot be eliminated. These defects can be plugged by coating a thin layer of a highly
permeable polymer such as silicone rubber or high-permeability fluoropolymer on
the surface of the thin, dense, nonporous, selective skin layer of the membrane.
Preparation of asymmetric integrally skinned polymeric hollow-fiber membranes
via the phase-inversion process is similar to that of the asymmetric integrally
skinned flat-sheet membranes. One major difference is that a casting knife is
used to cast a thin layer of polymer solution, which will then be converted into
asymmetric integrally skinned flat-sheet membrane, but an annular spinneret is
used to extrude a hollow-fiber polymer solution, which will then be converted into
asymmetric integrally skinned hollow-fiber membrane. Another major difference
is that a highly porous substrate such as a polyester or nylon cloth, which will
provide the membrane mechanical strength, is normally used for the fabrication of
asymmetric integrally skinned flat-sheet membranes, but preparation of polymeric
hollow-fiber membranes does not need any substrate and the membranes are
freestanding membranes. In addition, normally the polymer spinning solution
for the preparation of hollow-fiber membranes has higher polymer concentration
and therefore higher solution viscosity than the polymer-casting solution for the
preparation of flat-sheet membranes. The composition of the polymer-casting
9.2 Types and Preparation of Membranes 211
solution or spinning solution and the coagulation media are two key parameters
that determine the final membrane separation performance.
Overall, the thin, dense, nonporous, selective skin layer of the asymmetric
integrally skinned polymeric membranes will not be delaminated easily from
the thick, porous support layer because both layers are formed simultaneously
from a one-step phase-inversion process. However, fabrication of defect-free, high
selectivity asymmetric integrally skinned polymeric membranes sometimes is
challenging. The presence of nanopores or defects in the selective skin layer reduces
the membrane selectivity. In some cases, the high shrinkage of a high-performance
new polymeric membrane on the highly porous substrate during membrane casting
and drying process results in unsuccessful fabrication of asymmetric integrally
skinned flat-sheet membranes using the phase-inversion process. In some other
cases, novel polymeric materials have very limited solubility in common organic
solvents that are used for the fabrication of asymmetric integrally skinned polymeric
membranes, preventing the development of high-performance membranes.
dense, nonporous or almost nonporous, selective skin layer and the thick, porous
support layer are formed separately in two steps from different polymeric materials.
Similar to asymmetric integrally skinned membranes, the thin, dense, nonporous or
almost nonporous, selective skin layer is responsible for resistance and selectivity. It
is worth noting that the surface pore size control for the asymmetric porous support
membrane that will provide the TFC membrane mechanical strength and durability
will contribute to the TFC membrane separation performance even though the
asymmetric porous support membrane itself has no separation selectivity. The
two-step method for the preparation of bilayer TFC polymeric membranes typically
includes the fabrication of the asymmetric integrally skinned porous support poly-
meric membrane via the phase-inversion process followed by adding a thin selective
layer formed from a different polymer on the surface of the support membrane by a
coating method such as dip coating, spray coating, spin coating, laminating, inter-
facial polymerization, knife casting, or other methods. Some of the TFC polymeric
membranes have trilayers with an intermediate layer (or so-called gutter layer)
between the top thin selective skin layer and the bottom thick, porous, nonselective
support layer (Figure 9.3) to make a defect-free continuous selective skin layer [12].
Since the thin selective skin layer is formed separately from the thick, porous,
nonselective support layer in the TFC polymeric membranes, the thin selective
skin layer can be delaminated from the thick, porous, nonselective support layer in
some cases, which will result in significantly decreased selectivity for separations.
In addition, the porous, nonselective support layer needs to be insoluble in the
solvents that will be used for the preparation of the coating solution to form the
thin selective skin layer. Otherwise, the porous, nonselective support layer will
be damaged during the formation of the thin selective skin layer via the normal
coating process. To avoid these issues, TFC polymeric membranes have been
prepared via a one-step co-casting [13] or co-extruding [14] method in recent
years. TFC polymeric flat-sheet membranes can be fabricated by simultaneously
co-casting two different polymeric casting solutions containing the same polymer
or different polymers using two casting knives via a one-step phase-inversion
process. One solution at the bottom layer forms the low-cost, asymmetric, integrally
skinned, porous, nonselective bottom support layer and the other solution forms
another asymmetric integrally skinned high-performance selective layer on top
of the support layer. TFC polymeric hollow-fiber membranes can be fabricated
via a one-step co-extrusion phase-inversion process using two different polymeric
spinning solutions and a triple-annular spinneret. The core solution forms the
low-cost, asymmetric, integrally skinned, porous, nonselective core layer, and
the sheath solution forms the asymmetric integrally skinned high-performance
9.2 Types and Preparation of Membranes 213
selective sheath layer. The bore fluid normally comprising an organic solvent and
water creates the center hollow of the hollow-fiber membrane.
Preparation of TFC polymeric membranes using one-step co-casting or
co-extruding approach allows the use of higher-cost, high-performance poly-
mer to form the asymmetric integrally skinned selective layer and the use of
inexpensive, commercially available polymer to form the asymmetric, integrally
skinned, porous, nonselective support layer without any potential delamination
issue and the issue of support membrane damage.
Similar to asymmetric integrally skinned polymeric membranes, TFC polymeric
membrane may also comprise a thin coating layer of a highly permeable polymer
such as silicone rubber or high-permeability fluoropolymer on the surface of the
thin, dense, selective skin layer of the membrane to plug the defects and further
improve the membrane selectivity for gas and vapor separation applications.
The development of new polymeric membranes from the starting polymers to
the large-scale commercial applications involves many critical steps. Not only poly-
meric materials, but also membrane engineering need to be studied in detail. The
selection or synthesis of the polymeric membrane material for a specific separa-
tion application is a key initial step. Membrane engineering, including membrane
fabrication, element/module preparation, process design, and simulation, is equally
important or even more important than membrane material study. Polymeric casting
or spinning solution study allows the successful conversion of a polymeric material
to a commercially viable membrane with either flat-sheet or hollow-fiber geome-
try and each of them can have an asymmetric integrally skinned or TFC structure.
The membrane element/module preparation involves not only advanced polymeric
membrane, but also unique design of the element/module such as permeate and feed
spacer for spiral wound element, as well as the glue that holds membrane leaves or
hollow fibers together to prevent leaks. To extract the best value from the membrane,
sometimes membrane post-treatment such as chemical or physical cross-linking is
used to further enhance the membrane performance. Membrane performance eval-
uation and technoeconomic analysis allow further membrane optimization. Overall,
polymeric material selection, design, and synthesis would be less than about 20% of
work in the whole membrane development process.
Consistent, long-term membrane research programs in academia and R&D
platforms in industry will continuously advance membrane technologies and lead
to breakthrough membranes for separations. These activities should include novel
membranes such as molecular design and synthesis of polymers, novel membrane
components, improved element/module materials, and application-guided concept
discovery, new membrane processes such as physical/computational modeling,
process intensification, invention of new membrane-based processes, as well as
advanced membrane characterization techniques.
(COF), carbon molecular sieve, ceramic, silica, metal, and stainless-steel membranes
have several advantages over polymeric membranes for a wide range of applications.
Inorganic membranes have high thermal stability, high chemical resistance and
resistance to harsh environments, high permeability, conductivity, and selectivity
for some applications. Therefore, inorganic membranes provide new separation
opportunities such as high-temperature catalytic membrane reactor applications.
Significant research was reported on four main types of inorganic membranes based
on their preparation methods in the early 1990s, including sol–gel membranes such
as alumina, titanium, zirconia, and silica membranes; chemical vapor deposition
(CVD) membranes such as silica, zirconia, and palladium membranes; pyrolysis
membranes such as carbon and carbon molecular sieve membranes; as well as
hydrothermal membranes, such as microporous crystalline zeolite membranes.
Many different microporous crystalline MOF and ZIF hydrothermal membranes
such as MOF-5, ZIF-8, and ZIF-67 have been studied in academia since the report
of the first MOF-5 membrane for gas separation in 2009 [15]. This chapter will focus
on the discussion of microporous crystalline zeolite, MOF, and ZIF membranes for
separations based on molecular sieving separation mechanism.
B A
Polymer matrix
MMM
Robeson’s 1991
polymer upper bound
AlPO-14/P MMMs
𝛼CO2/CH4
P-3
P-2 P-1
PCO
2
~100 μm
same mixed-matrix material comprising dispersed inorganic particles and the con-
tinuous polymer matrix for the asymmetric integrally skinned MMMs (Figure 9.7).
TFC MMMs comprise a thin mixed-matrix selective layer on a highly porous
nonselective support layer, which can be formed from a polymeric or inorganic
material different from the mixed-matrix material in the selective layer. For asym-
metric MMMs to achieve improved selectivity and permeance compared to the
asymmetric polymeric membranes, the formation of a thin (<∼100 nm), defect-free
mixed-matrix selective layer is critical.
Most of the asymmetric MMMs that have been reported so far are asymmetric inte-
grally skinned MMMs with either flat-sheet or hollow-fiber geometry prepared from
mixed-matrix casting or spinning dopes via the phase-inversion process [56–59]. The
compositions of the mixed-matrix casting or spinning dopes are different from those
for mixed-matrix dense films and normally comprise several different organic sol-
vents and nonsolvents as well as some additives to control the mixed-matrix selective
layer thickness and the morphology of the MMMs. Preparation of a stable, uni-
form mixed-matrix casting or spinning dope with the inorganic particles uniformly
dispersed in the polymer solution without particle agglomeration is critical for the
formation of a defect-free mixed-matrix selective layer, which will provide the MMM
with enhanced selectivity. Some of the asymmetric MMMs are TFC MMMs prepared
33065-414(2)
Residual
Feed spacer
Membrane Residual
Permeate spacer
Membrane
Feed spacer
have been used commercially for RO, NF, UF, MF, and gas separations. There are
two types of hollow-fiber modules for these applications, including shell-side feed
modules and bore-side feed modules. Shell-side feed modules are generally used for
high-pressure applications up to about 1500 psig feed pressure such as for CO2 /CH4
separation and H2 purification with counter-current flow or cross-flow. Fouling on
the feed side of the membrane can be a problem with this kind of module design,
and pretreatment of the feed stream to remove particulates is required. Bore-side
feed modules are generally used for low- and medium-pressure feed streams up to
about 200 psig with counter-current flow, such as separation of N2 or water from
air, where good flow control to minimize fouling and concentration polarization on
the feed side of the membrane is desired.
Multichannel monolith module is the most preferred module configuration
for high-cost inorganic membranes such as the commercial ceramic membranes
and NaA zeolite membranes to achieve high membrane area packing density and
reduce membrane system cost for large-scale membrane separation processes with
high feed treating capacity and high product purity and recovery. The areas of the
inorganic membrane in each multichannel monolith module are about 10–15 m2 .
Several multichannel monolith zeolite membrane modules have been reported in
the literature [63, 64]. NaA zeolite membranes with multichannel monolith module
design have been used commercially in isopropanol and bioethanol dehydration
plants in China and India [25, 65].
Membrane technologies have been used for a wide range of large-scale industrial
gas, vapor, liquid, and ion separation processes such as RO, natural gas upgrad-
ing, hydrogen recovery, organic vapor/gas and vapor/vapor separations, pervapora-
tion, NF, organic solvent NF, UF, MF, membrane contactor, electrodialysis, fuel cell,
and other electrochemical processing. Innovation in membrane materials, equip-
ment, and process technology integration has been instrumental in enabling the
widespread use of membranes in the chemical industry. Membrane technologies
have been widely adopted in semiconductor, automotive, water, food, pharmaceuti-
cal, biotechnology industries, and a wide range of environmental applications. Mem-
brane technology alone or its combination with other separation technologies in
integrated systems has been a successful approach for seawater desalination to meet
the freshwater demand in many regions of the world at lower cost and minimum
environmental impact. Worldwide sales of membrane products and systems have
seen a significant growth and increase in applications in the last two decades.
widely used technology for CO2 and H2 S removal from natural gas. Amine plants,
however, suffer from problems such as high capital cost, complicated operation,
and expensive maintenance. Membrane separation is an alternative technology
for this natural gas upgrading application. Both spiral wound and hollow-fiber
polymeric membranes have been adopted commercially for this separation. Some
subquality natural gas reserves have high N2 content, which needs to be removed
to meet natural gas pipeline specifications. Cryogenic distillation and pressure
swing adsorption (PSA) technologies are currently used for this separation. MTR
has commercialized NitroSepTM membrane, which is more permeable to methane,
ethane, and other hydrocarbons than to N2 , for separating N2 from natural gas [7].
UOP’s Separex membrane technology has been an important enabler for effective
and efficient separation of contaminants such as CO2 and H2 S from natural gas
for onshore and offshore applications. Separex membrane technology was first
commercialized in small membrane units of <10 MMSCFD (million standard
cubic feet per day) treating capacities in early 1980s and it was used when
the remote locations of natural gas reserves required a simple and reliable
separation technology. Separex membrane systems became larger and larger with
the gas-treating capacity increased to 1000 MMSCFD in 2009 (Figure 9.9) and
have been extended from onshore applications to offshore platforms and most
recently offshore FPSO vessels that require lower-weight and smaller-footprint
membrane systems (Table 9.1). Figure 9.10 shows a single-lift Separex module
that offers small footprint and low weight for FPSO vessels. The simplest Separex
membrane-processing scheme is a one-stage flow scheme. For the one-stage flow
scheme (Figure 9.11), a natural gas feed is separated into a hydrocarbon-rich,
low CO2 and H2 S residual product stream and a CO2 and H2 S-rich permeate
stream. High hydrocarbon recovery (>95%) can be achieved using a two-stage
(Figure 9.12) or multistage flow scheme where the low-pressure, first-stage
permeate is compressed and processed in a second-stage membrane. For
example, a two-stage Separex membrane system was used for a large offshore
platform in Malaysia (Figure 9.13) to process 680 MMSCFD of natural gas. The
system reduced CO2 from 44 to 8 mol%. Another Separex membrane-processing
scheme is a two-step flow scheme (Figure 9.14) where the residue from the first
membrane is sent to the second membrane to further remove CO2 and H2 S and
the permeate from the second membrane is compressed and sent back to the
first membrane feed. The two-step membrane system has better hydrocarbon
recovery than that of the one-stage system, but is worse than that of the two-stage
system.
Separex commercial membrane systems include not only the membrane unit, but
also the membrane pretreatment unit. Proper membrane pretreatment design
is critical to the performance of the membrane systems. Many of Separex mem-
brane pretreatment systems include UOP MemGuardTM pretreatment system
(Figure 9.15), which is a regenerative system that can provide an absolute cutoff
of heavy hydrocarbons and can remove water, mercury, and other contaminants
along with the heavy hydrocarbons. MemGuard pretreatment systems have been
used for both onshore and offshore membrane systems. The invention of new
230 9 Membranes
1000
100
Feed flow (MMSCFD)
10
1 Avg size
0.1
1980 1985 1990 1995 2000 2005 2010
Year
Figure 9.9 Separex commercial membrane systems for natural gas upgrading with
significantly increased gas treating capacities from 1981 to 2009.
Table 9.1 Separex commercial membrane systems for offshore CO2 removal from natural
gas applications.
Treating capacity
Location (MMSCFD) CO2 removal (%) Stages
Thailand 32 50 to 20 1
Thailand 480 35 to 23 1
Thailand 580 40 to 23 1
Malaysia 680 44 to 8 2
FPSO Thailand 11 28 to 3 1
FPSO Brazil (14 units) 176–240 10–40 to 3–5 1–2
Feed
Permeate
(CO2 and H2S enriched)
Membrane unit
highly porous support layer are also based on solution-diffusion separation mech-
anism. Rubbery polymers such as silicone rubber with high solubility selectivity
for vapor over N2 or other gases have been used for the development of commer-
cial membranes for vapor/gas separations. Vapor/gas separation membranes are
commercially available from MTR, Dalian Eurofilm Industrial Ltd., and BORSIG
Membrane Technology GmbH.
232 9 Membranes
Residue
Feed
Permeate
Optional
Regenerable
Chiller and Coalescing
Feed MemGuardTM
separator filter
system
Particle Membrane
Heater
filter
Sales gas
Figure 9.15 Separex membrane system, including MemGuard system and membrane unit.
Ligh pressure R
the low-pressure permeate side of the FTM. FTMs have been prepared in lab scale
and some of them have been evaluated in demo unit for propylene/propane sep-
aration [81–83]. Membrane performance stability needs to be further improved
and the issue of complexing agent poisoning by contaminants needs to be solved
for commercial applications of FTMs for olefin/paraffin separations.
Pervaporation: Pervaporation is a membrane separation process to separate liquid
mixtures that combines evaporation of the volatile liquid component with
permeation of the vapor-phase component through a pervaporation membrane.
Vacuum is normally applied to the permeate side of the membrane to facilitate
the separation and recover the permeate. Both polymeric and inorganic mem-
branes have been used for commercial-scale pervaporation processes to separate
azeotropic mixtures such as alcohol and organic solvent dehydration, but sales
of pervaporation systems are much lower than other membrane separation pro-
cesses. Pervaporation using polymeric membranes is based on solution-diffusion
separation mechanism same as that for polymeric gas separation membranes.
As with polymeric gas separation membranes, polymeric pervaporation mem-
branes prepared from polymers such as polyvinyl alcohol or silicone rubber
have a thin, nonporous, selective skin layer on the surface of a highly porous
support layer. Pervaporation using small-pore microporous inorganic zeolite
membranes such as NaA membrane is based on preferential adsorption and/or
molecular sieving separation mechanism and the membranes have very high
selectivity. Most of the pervaporation membranes are commercially available
from GKSS (polymeric membrane), MTR (polymeric membrane), Jiangsu Nine
Heaven (NaA), Mitsui-BNRI ZeoSepA (NaA), Mitsubishi KonkerTM (NaA), and
Mitsubishi ZEBREXTM (high Si/Al ratio CHA).
Nanofiltration (NF): NF membrane is a loose RO membrane with an average pore
size of 1–2 nm in the thin selective skin layer and has been widely used for water
softening and treatment of high color and high organic content feed water. NF
separation is based on sorption-sieving separation mechanism. NF is superior to
both RO and UF for the treatment of bleaching effluents from pulp and paper
manufacturing plants due to its higher water flux than RO and higher rejection of
bleaching agent than UF. NF membrane can be operated at lower pressure than
RO membrane and has high rejection for divalent ions (>95%) and organics, but
has lower rejection for NaCl (10–50%) than RO membrane. Most of the commer-
cially available NF membranes are PA TFC membranes offered by Dow FILMTEC,
hydranautics, and several other companies. OSNF membranes are also loose RO
type of membranes with an average pore size of 1–2 nm in the thin selective skin
234 9 Membranes
layer and have high resistance to organic solvents. The most common polymers
for OSNF membranes are polyimides. OSNF-based MAX-DEMAX process devel-
oped by W. R. Grace and ExxonMobil has been used commercially for crude oil
dewaxing. OSNF is still an emerging membrane technology and there have been
very limited large-scale commercial sales for OSNF systems.
Ultrafiltration (UF): UF membrane is a porous membrane with an average pore size
of 1–100 nm in the surface skin layer and has been widely used for surface water
treatment, RO pretreatment, colloidal silica and iron removal from water, color
reduction, oil/water separation, milk processing, fruit juice clarification, automo-
bile paint recovery, biomolecule separation, as well as therapeutic proteins, DNA,
and RNA purification and filtration. UF membrane separation is based on not
only the pore size of the membrane, but also the shape and size of the molecules
to be rejected, as well as the solute/membrane interactions. The most common
UF membranes are prepared from CA, PSf, PES, polyacrylonitrile (PAN), PVDF,
PEI, or ceramic inorganic materials. Most recently, several approaches have been
studied to enhance the flux, reduce fouling, and improve solvent resistance of UF
membranes such as polymer chain modification, use of solvent-resistant polymers
such as PEEK and PBI, addition of hydrophilic polymer, blending of two different
polymers, modification of membrane surface, and coating of ceramic membrane
with silica or zirconia.
Microfiltration (MF): MF membrane is another porous membrane with an average
pore size of 0.1–10 μm in the surface skin layer and has been widely used for bio-
engineering to remove microorganisms from the fermentation products or remove
yeast from alcoholic beverages, removing particulates, colloidal materials, and
complexed heavy metals from liquid suspensions, and cell harvesting. MF mem-
brane separation is more complicated than simple sieving due to the pore size
reduction by the particles or macromolecules to be separated. The most common
MF membranes are prepared from CTA, PSf, PVDF, PP, polyethylene (PE), poly-
carbonate, polyester, PI, polytetrafluoroethylene (PTFE), nylon 6, nylon 6,6, or
ceramic inorganic materials.
UF and MF membranes are commercially available from a significant number
of companies such as Hydranautics, Suez, Hyflux, Pall, Tianjin MOTIMO
Membrane Technology, Litree Purifying Technology, Beijing Scinor Membrane
Technology, Beijing OriginWater Technology, Mitsubishi, and Jiangsu Jiuwu
Hi-Tech.
Membrane contactor: Membrane contactor is a membrane-based device that is used
to achieve gas/liquid mass transfer by allowing a gaseous phase and a liquid
phase to come into direct contact with each other without dispersion of one phase
within another. In a membrane contactor, the membrane separation is integrated
with a conventional phase contacting operation such as absorption or extraction
to exploit the benefits of both technologies. The key features for membrane
contactors are nonsensitivity to flooding, channeling, or back-mixing, modular
design, much higher surface area per unit volume as compared to conventional
columns or reactors, and use of vacuum and inert gas to create the driving force
for mass transfer. Common membranes used for membrane contactors are UF
References 235
References
24 Kim, S.J., Liu, Y., Moore, J.S. et al. (2016). Thin hydrogen-selective SAPO-34 zeo-
lite membranes for enhanced conversion and selectivity in propane dehydrogena-
tion membrane reactors. Chemistry of Materials 28: 4397.
25 Rangnekar, N., Mittal, N., Elyassi, B. et al. (2015). Zeolite membranes-a review
and comparison with MOFs. Chemical Society Reviews 44: 7128.
26 Nakayama, K., Suzuki, K., Yoshida, M., et al. (2006). Method for preparing DDR
type zeolite membrane, DDR type zeolite membrane, and composite DDR type
zeolite membrane, and method for preparation thereof. US Patent 7,014,680.
Issued: 21 March 2006.
27 Uchikawa, T., Yajima, K., Nonaka, H., and Tomita, T. (2013). Method for produc-
tion of DDR type zeolite membrane. US Patent 8,377,838. Issued: 19 February
2013.
28 Teranishi, M., Miyahara, M., Ichikawa, M., and Suzuki, H. (2016). Porous body
and honeycomb-shaped ceramic separation-membrane structure. US Patent
9,321,016. Issued: 26 April 2016.
29 Imasaka, S., Masaya, I., Asegawa, Y., et al. (2016). Zeolite membrane, production
method therefor, and separation method using same. WO Patent Application
2016/006564. Published on 14 January 2016.
30 Tang, Z., Kim, S.J., Gu, X., and Dong, J. (2009). Microwave synthesis of MFI-type
zeolite membranes by seeded secondary growth without the use of organic struc-
ture directing agents. Microporous and Mesoporous Materials 118: 224.
31 Shi, H. (2015). Organic template-free synthesis of SAPO-34 molecular sieve mem-
branes for CO2 –CH4 separation. RSC Advances 5: 38330.
32 Agrawal, K.V., Topuz, B., Pham, T.C.T. et al. (2015). Oriented MFI membranes by
gel-less secondary growth of sub-100 nm MFI-nanosheet seed layers. Advanced
Materials 27: 3243.
33 Truter, L.A., Ordomsky, V.V., Nijhuis, T.A., and Schouten, J.C. (2012).
Preparation of ZSM-5 zeolite coatings within capillary microchannels.
Journal of Materials Chemistry 22: 15976.
34 Kwon, H.T. and Jeong, H.K. (2013). In situ synthesis of thin zeolitic-imidazolate
framework ZIF-8 membranes exhibiting exceptionally high propylene/propane
separation J. American Chemical Society 135: 10763.
35 Bein, T. (1996). Synthesis and applications of molecular sieve layers and mem-
branes. Chemistry of Materials 8: 1636.
36 Lai, Z., Bonilla, G., Diaz, I. et al. (2003). Microstructural optimization of a zeolite
membrane for organic vapor separation. Science 300: 456.
37 Pham, T.C.T., Nguyen, T.H., and Yoon, K.B. (2013). Gel-free secondary growth of
uniformly oriented silica MFI zeolite films and application for xylene separation.
Angewandte Chemie, International Edition 52: 8693.
38 Pham, T.C.T., Kim, H.S., and Yoon, K.B. (2011). Growth of uniformly oriented
silica MFI and BEA zeolite films on substrates. Science 334: 1533.
39 Kwon, H.T., Jeong, H.K., Lee, A.S. et al. (2015). Heteroepitaxially grown zeolitic
imidazolate framework membranes with unprecedented propylene/propane
separation performances. Journal of the American Chemical Society 137: 12304.
238 9 Membranes
40 Choi, J., Jeong, H.K., Snyder, M.A. et al. (2009). Grain boundary elimination in a
zeolite membrane by rapid thermal processing. Science 325: 590.
41 Schillo, M.C., Park, I.S., Chiu, W.V., and Verweij, H. (2010). Rapid thermal
processing of inorganic membranes. Journal of Membrane Science 362: 127.
42 Huang, Y., Wang, L., Song, Z. et al. (2015). Growth of high-quality,
thickness-reduced zeolite membranes towards N2 /CH4 separation using
high-aspect-ratio seeds. Angewandte Chemie, International Edition 54: 10843.
43 Varoon, K., Zhang, X., Elyassi, B. et al. (2011). Dispersible exfoliated zeolite
nanosheets and their application as a selective membrane. Science 334: 72.
44 Zhang, H., Xiao, Q., Guo, X. et al. (2016). Open-pore two-dimensional MFI zeo-
lite nanosheets for the fabrication of hydrocarbon-isomer-selective membranes
on porous polymer supports. Angewandte Chemie, International Edition 55: 7184.
45 Jeon, M.Y., Kim, D., Kumar, P. et al. (2017). Ultra-selective high-flux membranes
from directly synthesized zeolite nanosheets. Nature 543: 690.
46 Peng, Y., Li, Y., Ban, Y. et al. (2014). Metal-organic framework nanosheets as
building blocks for molecular sieving membranes. Science 346: 1356.
47 Morigami, Y., Kondo, M., Abe, J. et al. (2001). The first large-scale pervaporation
plant using tubular-type module with zeolite NaA membrane. Separation and
Purification Technology 25: 251.
48 Mahajan, R. and Koros, W.J. (2000). Factors controlling successful formation
of mixed-matrix gas separation materials. Industrial and Engineering Chemistry
Research 39: 2692.
49 Liu, C., Chiou, J.J., Wilson, S.T. (2009). Cross-linkable and cross-linked mixed
matrix membranes and methods of making the same. US Patent 7,485,173.
Issued: 3 February 2009.
50 Liu, C., Tang, M. W., Wilson, S. T., Lesch, D. A. (2010). Method of making
high performance mixed matrix membranes using suspensions containing poly-
mers and polymers stabilized molecular sieves. US Patent 7,815,712. Issued: 19
October 2010.
51 Liu, G., Chernikova, V., Liu, Y. et al. (2018). Mixed matrix formulations with
MOF molecular sieving for key energy-intensive separations. Nature Materials 17:
283.
52 Japip, S., Liao, K.S., and Chung, T.S. (2017). Molecularly tuned free volume of
vapor cross-linked 6FDA-durene/ZIF-71 MMMs for H2 /CO2 separation at 150 ∘ C.
Advanced Materials 29: 1603833.
53 Miller, S.J., Kuperman, A., Vu, D.Q. (2007). Mixed matrix membranes with small
pore molecular sieves and methods for making and using these membranes.
US Patent 7,166,146 B2. Issued: 23 January 2007.
54 Kulkarni, S.S., David, H.J., Corbin, D.R., Patel, A.N. (2003). Gas separation
membranes with organosilicone-treated molecular sieve. US Patent 6,508,860.
Issued: 21 January 2003.
55 Marand, E., Pechar, T.W., Tsapatsis, M. (2006). Mixed matrix membranes.
US Patent 7,109,140 B2. Issued: 19 September 2006.
References 239
56 Kulkarni, S.S. and Hasse, D.J. (2005). Novel polyimide based mixed matrix
membranes. US Patent Application 2005/0268782 A1. Published on 8 December
2005.
57 Ekiner, O.M. and Kulkarni, S.S. (2003). Process for making mixed matrix hollow
fiber membranes for gas separation. US Patent 6,663,805. Issued: 16 December
2003.
58 Jiang, L.Y., Chung, T.S., and Kulprathipanja, S. (2006). An investigation to
revitalize the separation performance of hollow fibers with a thin mixed matrix
composite skin for gas separation. Journal of Membrane Science 276: 113.
59 Ismail, A.F., Kusworo, T.D., and Mustafa, A. (2008). Enhanced gas permeation
performance of polyethersulfone mixed matrix hollow fiber membranes using
novel Dynasylan Ameo silane agent. Journal of Membrane Science 319: 306.
60 Jia, M., Peinemann, K.V., and Behling, R.D. (1992). Preparation and character-
ization of thin-film zeolite-PDMS composite membranes. Journal of Membrane
Science 73: 119.
61 Jeong, B.-H., Hoeka, E.M.V., Yan, Y. et al. (2007). Interfacial polymerization
of thin film nanocomposites: a new concept for reverse osmosis membranes.
Journal of Membrane Science 294: 1.
62 Hess, S.C., Grass, R.N., and Stark, W.J. (2016). MOF channels within porous
polymer film: flexible, self-Supporting ZIF-8 poly(ether sulfone) composite mem-
brane. Chemistry of Materials 28: 7638.
63 Liu, W., Post, P., Williams, J. L., et al. (2007). Multi-channel cross-flow porous
device. US Patent 7,169,213. Issued: 30 January 2007.
64 Fekety, C. R., Kinney, L. D., Liu, W., Song, Z. (2009). Zeolite membrane
structures and methods of making zeolite membrane structures. WO Patent
Application 2009005745A1. Issued: 8 January 2009.
65 Caro, J. and Noack, M. (2009). Zeolite membranes – status and prospective.
In: Advances in Nanoporous Materials, vol. 1 (ed. S. Ernst), 1. Elsevier.
66 Michaels, A.S. (1968). Separation techniques for the CPI (chemical process
industry). Chemical Engineering Progress 64: 31.
67 Liu, C., Xu, Y., Liao, S. et al. (1997). Selective hydrogenation of propadiene and
propyne in propene with catalytic polymeric hollow-fiber reactor. Journal of
Membrane Science 137: 139.
68 Liu, C., Xu, Y., Liao, S., and Yu, D. (1998). Mono- and bimetallic catalytic hollow
fiber reactors for the selective hydrogenation of butadiene in 1-butene. Applied
Catalysis A: General 172: 23.
69 Brandao, L., Fritsch, D., Mendes, A.M., and Madeira, L.M. (2007). Propylene
hydrogenation in a continuous polymeric catalytic membrane reactor. Industrial
and Engineering Chemistry Research 46: 5278.
70 Kita, H., Tanaka, K., Okamoto, K.I., and Yamamoto, M. (1987). The Esterifica-
tion of oleic acid with ethanol accompanied by membrane separation. Chemistry
Letters 16: 2053.
71 Solovieva, A.B., Belkina, N.V., and Vorobiev, A.V. (1996). Catalytic process of
alcohol oxidation with target product pervaporation. Journal of Membrane Sci-
ence 110: 253.
240 9 Membranes
72 Binning, R.C. and Kelly, J.T. (1959). Hydrocarbon conversion with dialytic sep-
aration of the catalyst from the hydrocarbon products. US Patent 2,913,507.
Issued: 17 November 1959.
73 Champagnie, A.M., Tsotsis, T.T., Minet, R.G., and Webster, A.I. (1990). A high
temperature catalytic membrane reactor for ethane dehydrogenation. Chemical
Engineering Science 45: 2423.
74 Lafarga, D., Santamaria, J., and Menendez, M. (1994). Methane oxidative
coupling using porous ceramic membrane reactors - I. reactor development.
Chemical Engineering Science 49: 2005.
75 Lu, Y., Dixon, A.G., Moser, W.R. et al. (2000). Oxidative coupling of methane
using oxygen-permeable dense membrane reactors. Catalysis Today 56: 297.
76 Capinelli, G., Carosini, E., Cavani, F. et al. (1996). Comparison of the catalytic
performance of V2 O5 /γ-Al2 O3 in the oxidehydrogenation of propane to propy-
lene in different reactor configurations: (i) packed-bed reactor, (ii) monolith-like
reactor and (iii) catalytic membrane reactor. Chemical Engineering Science 51:
1817.
77 Torres, M., Sanchez, J., Dalmon, J.A. et al. (1994). Modeling and simulation
of a three-phase catalytic membrane reactor for nitrobenzene hydrogenation.
Industrial and Engineering Chemistry Research 33: 2421.
78 Udovich, C.A. (1998). Ceramic membrane reactor for the conversion of natu-
ral gas to syngas. In: Natural Gas Conversion V, Studies in Surface Science and
Catalysis, vol. 119 (eds. A. Parmaliana, D. Sanfilippo, F. Frusteri, et al.), 417.
79 van Veen, H.M., Bracht, M., Hamoen, E., and Alderliesten, P.T. (1996). Funda-
mentals of Inorganic Membrane Science and Technology (eds. A.J. Burggraaf and
L. Cot), 641.
80 Baker, R.W. (2002). Future directions of membrane gas separation technology.
Industrial and Engineering Chemistry Research 41: 1393.
81 Pinnau, I., Tpy, L.G., and Casillas, C. (1997). Olefin separation membrane and
process. US Patent 5,670,051. Issued: 23 September 1997.
82 Herrera, P.S., Feng, X., Payzant, J.D., and Kim, J.-H. (2008). Process for the sepa-
ration of olefins from paraffins using membranes. US Patent 7,361,800. Issued: 22
April 2008.
83 Feiring, A.E., Laareri, J., Majumdar, S., and Shsnggusn N. (2018). Membrane
separation of olefin and paraffin mixtures. US Patent 10,029,248. Issued: 24 July
2018.
84 Ran, J., Wu, L., He, Y. et al. (2017). Ion exchange membranes: new developments
and applications. Journal of Membrane Science 522: 267.
241
10
Absorption
35
30
25
20
15
10
0
1950 1960 1970 1980 1990 2000 2010 2020
Figure 10.1 US Natural Gas Consumption (1950–2018) in trillion cubic feet. Source: US
Energy Information Administration [3].
as well as around the world (Figure 10.1). Natural gas production has risen steadily
over the past 50 years. As a result, the importance of absorption methods to the oil
and gas and chemical industries has been, likewise, increasing steadily. Proven natu-
ral gas reserves are around 190 trillion m3 globally, but total reserves are much larger
than this figure [4]. The Middle East and Russia hold the largest total natural gas
reserves, while the largest current national producer and consumer of natural gas is
the United States.
The main constituent of natural gas is methane (Table 10.2), but it can contain a
range of other higher hydrocarbons. While these hydrocarbons represent the main
attraction of natural gas extraction due to their energy value through combustion,
there also are many non-hydrocarbon materials present in natural gas such as H2 S
10.2 Acid Gas Removal 243
and CO2 , which can interfere with the transport and storage of natural gas and also
can damage expensive energy infrastructure if not properly treated via methods such
as chemical and physical absorption. In addition, the presence of CO2 can lead to a
reduction in energy value of the natural gas as there is no energy value unlocked
with the attempted combustion of this contaminant.
Given its importance in the energy market, the application of gas separation meth-
ods to the removal of acidic contaminants such as H2 S and CO2 from natural gas will
be the focus of this chapter. It is important to note that many of these same chemi-
cal and physical absorption methods also are effective in the treatment of synthesis
gas (CO + H2 ), as both H2 S and CO2 require sweetening for the same reasons as
natural gas.
Gas absorption can be described as the removal of one or more pollutants from a
contaminated gas stream via intimate contact of that contaminated gas with a liq-
uid (absorber) that enables the pollutants to dissolve within the absorption liquid.
The principal factors dictating absorption performance are the solubility of the con-
taminants in the absorbing liquid as well as the ability of the absorption units to
efficiently mix gas and absorption liquid through the creation of a large amount of
interfacial surface area.
The corrosion chemistry of H2 S and CO2 is described in the equations below. Mild
grades of steel (e.g. carbon steel) often are pitted through the action of acid gas con-
taminants. Removal of these materials via chemical or physical absorption helps
preserve energy infrastructure, including wells and pipelines. In addition to H2 S and
CO2 levels, corrosion also is influenced by temperature, water chemistry (esp. pH),
flow velocity, oil or water wetting, and composition and the surface condition of the
steel.
H 2 S corrosion reactions:
Conventional processes for removing acid gases typically involve their countercur-
rent absorption from the contaminated natural gas or syngas stream using a regen-
erative solvent in an absorber column. This approach of contacting a gas stream to a
liquid absorber for the removal of acid gases is common across a wide range of pro-
cess industries, including natural gas production, refining, and chemical production.
Because each of these industries requires different degrees of acid gas removal, the
absorption solvent employed will vary significantly.
Shown in Table 10.3 are some common solvents (both chemical and physical)
employed in acid gas absorption/separation. The criteria for solvent selection in
chemical and physical absorption will be covered in more detail later in this chapter.
For now, please note that primary, secondary, and tertiary amines all find use in
chemical absorption processes. In addition, non-amine systems are used (e.g. potas-
sium carbonate in the removal of CO2 via the BenfieldTM process).
Prior to selecting either a chemical of physical absorption approach, the strengths
and weaknesses of each need to be considered. While the loading of a physical sol-
vent is proportional to the partial pressure of a component in natural gas, the loading
of a chemical solvent is determined by the stoichiometric reaction between a com-
ponent in natural gas and chemical solvent (i.e. it is a weak function of pressure).
At low partial pressure of the component (generally low concentration), a chemical
10.3 Chemical Solvent 245
Chemical Physical
1
Physical solvent
Chemical solvent
Partial pressure
As demonstrated in 10.2, when impurity concentrations are low (e.g. 1–10% H2 S and
CO2 ), chemical absorption is favored over physical absorption. The ideal solvent
246 10 Absorption
Table 10.4 Qualitative characteristics of the amines for the removal of H2 S and CO2 .
for this process is capable of a fast, reversible chemical reaction with the gaseous
impurity. When impurity levels are even lower, solvents that react irreversibly are
employed, but such irreversible systems bring drawbacks such as solid waste treat-
ment and added cost. Chemical solvents include aqueous solutions of primary, sec-
ondary, and tertiary amines (see Table 10.4). We will separately consider inorganic,
alkaline salts such as potassium carbonate (K2 CO3 ) in our summary of the Benfield
process below.
Through acid-base reactions, aqueous solutions of these basic compounds capture
and remove acid gases and other impurities by forming weak chemical bonds with
dissolved acid gases (H2 S and CO2 ) in the absorber. The bonds are broken by heat in
the regenerator to release the acid gases and regenerate the solvent for reuse.
1. The process of separating an acidic gas from gaseous mixtures, which includes
effecting intimate contact of a gaseous mixture with an absorbent agent in liquid
form including an amine selected from the group consisting of aliphatic and
cycloparaffin amines, and which is free from carboxyl or carbonyl groups, and
which has a boiling point not substantially below 100 ∘ C. – Bottoms [8]
The basic design of a chemical absorption unit is included in Figure 10.3. The flow
of gas upward and liquid amine downward encourages the sort of intimate mixing
required of a high efficiency chemical (or physical) absorption process. The packing
material within the column bed plays a role in creating the mixing needed to effi-
ciently remove acid gas impurities from natural gas and other hydrocarbon streams.
10.3 Chemical Solvent 247
Liquid Gas
in out
Packed bed
Liquid Gas
out in
(1) Required purity of cleaned gas stream for further processing or end use
(2) Composition of feed gas
(3) Utility requirements, process costs
(4) Corrosion and solvent degradation/solvent losses.
We will now dive a bit deeper into the strengths of each main category of amine
solvent for chemical absorption.
Table 10.5 Four typical AGFS (amine guard formulated solvent) flow schemes.
Reboiler duty
Desired product Typical (MBtu/lbmol
Scheme Feed gas quality gas quality application CO2 removed)
Flash only Very high acid gas >2% CO2 Pipeline NG 8–10
(>12%)
Conventional Low acid gas (<7%) 50 ppm CO2 LNG 45–60
1-stage High acid gas 50–1000 ppm CO2 LNG, NGL 32–40
(7–12%)
2-stage Very high acid gas 500 ppm CO2 Ammonia 12–18
(>12%)
low corrosivity. Oxidation of MDEA forms corrosive acids, which can result in the
buildup of iron sulfide in the absorption system. [10].
MDEA offers several unique advantages over primary and secondary amines,
including lower vapor pressure, lower heat of reaction, higher resistance to degra-
dation, lower corrosivity, and higher selectivity toward H2 S in the presence of CO2 ;
being a weaker base, MDEA reacts much more quickly with H2 S than CO2 . Its
lower heat of reaction allows MDEA to be used in pressure swing units for bulk
CO2 removal. In a pressure swing plant, the rich amine is flashed at or around
atmospheric pressure and little or no heat is required for stripping.
Once the correct amine for chemical absorption has been selected, the design of
the flow scheme must be considered. There are four basic flow scheme options for
an amine guard formulated solvent (AGFS) chemical absorption unit (Table 10.5).
The first scheme (flash) is a simple flow scheme and is relatively inexpensive, has
a low energy requirement, is ideal for bulk removal of CO2 and partial removal of
H2 S to generate pipeline quality natural gas. The second flow scheme (conventional,
depicted in Figure 10.4) is suitable for both LNG and natural gas liquids (NGL); it can
achieve CO2 levels below 50 ppm, H2 S levels below 4 ppm, and requires less solvent
than the flash system. The third option (1-stage) is capable of the same CO2 and H2 S
removal as a conventional unit but at lower energy usage. Finally, the 2-stage system
is suitable for LNG and ammonia and provides the flexibility to tradeoff solvent flow
rate for thermal regenerator duty (at much lower regeneration duty requirements
than the 1-stage system).
Acid gas
Acid gas
Sweet KO drum
gas
Make-Up
Water
Acid gas
cooler
Amine Lean Lean
absorber solution solution Reflux
pump cooler pump
Flash gas
to fuel
header Filtration
Amine
system Make-Up
Feed Make-up stripper
Water
gas water
Rich
flash drum
Amine
roboiler
Lean/rich
Lean solution
exchanger
booster pump
Figure 10.4 Honeywell UOP conventional chemical solvent – amine process (suitable for
LNG and NGL).
does not provide acceptable treated gas purity. Caustic solutions, either NaOH or
KOH, could remove the acid gases very effectively, but could not be regenerated.
Using a solution of KOH to first pick up CO2 would generate potassium carbon-
ate in solution, which could absorb still more acid gases. Thus, the “hot potassium
carbonate” process was born.
The UOP Benfield Process is a thermally regenerated cyclical solvent process using
an activated, inhibited hot potassium carbonate solution to remove CO2 , H2 S, and
other acid gas components. More than 700 Benfield units have been put into com-
mercial service. In addition to its wide use in ammonia and hydrogen production,
the Benfield process has been applied in 50+ natural gas plants and in direct iron
ore reduction plants.
The absorption of CO2 into an aqueous potassium carbonate (K2 CO3 ) system
means there are two reaction paths available. Either hydroxide ions react with CO2
to directly form bicarbonate ions:
or water can hydrate CO2 to form carbonic acid, which subsequently deprotonates
to form bicarbonate:
Acid gas
Purified gas
Feed gas
Absorber
Regeneration
Figure 10.5 BenfieldTM process: chemical absorption based on K2 CO3 (conventional flow
scheme).
Shown in Figure 10.5 is the conventional flow scheme for the Benfield process,
including the thermal regeneration system. With over 700 units in commercial
operation, the Benfield process offers efficiency coupled with low capital invest-
ment. This method of separation achieves very low CO2 concentrations (<50 ppm)
in treated gas as well as low H2 S levels (<4 ppm)
As shown in Figure 10.2, physical solvent absorption systems offer the advantage of
operating at high loading. Indeed, solvent loading increases linearly with the partial
pressure of the contaminant to be removed. Therefore, when the component to be
removed (e.g. CO2 or H2 S) is present at a high partial pressure, physical absorption
252 10 Absorption
10.4.1.1 DMEPG (Dimethyl Ether of Polyethylene Glycol, (CH3 O(C2 H4 O)n CH3 ))
DMEPG is used in UOP’s Selexol process and is a mixture of dimethyl ethers of
polyethylene glycol where n falls in the range of two to nine. DMPEG is often used to
physically absorb H2 S, CO2 , and mercaptans from gas streams. As can be seen from
the data in Table 10.7, DMEPG can be used for selective H2 S removal, which requires
stripping, vacuum stripping, or a reboiler. Depending on downstream unit configu-
ration, the H2 S-rich feed can be sent onward to the Claus unit. It also is possible to use
DMEPG for selective H2 S removal with deep CO2 removal typically accomplished
via a two-stage process with two absorption and regeneration columns. The first
column selectively removes H2 S, while CO2 is removed in the second absorber. The
second-stage solvent can be regenerated with air or nitrogen for deep CO2 removal,
or using a series of flashes if bulk CO2 removal is required.
Compared to the other two solvents (methanol and propylene carbonate), DMEPG
is of higher viscosity. This difference in rheology reduces mass transfer rates and tray
efficiencies and increases packing (or tray number) requirements. These differences
will become even more important when operating at reduced temperatures. DMEPG
service temperatures range from −18 ∘ C up to 175 ∘ C.
For each project, a specific design is required to achieve the desired product qual-
ity and the optimum acid gas quality at the same time. There are three general
flow schemes for physical absorption, which differ primarily in their regeneration
approach (Table 10.8):
A flash regeneration system involved the enriched solvent being regenerated solely
through pressure let-down; there is no thermal regeneration involved. This method
is best suited for bulk CO2 removal to meet pipeline specs (<3% CO2 ) and is not
applicable for meeting low CO2 (e.g. treated gas) specifications or for high sulfur
Treated
gas
CO2
absorber
CO2 Acid
gas
Stripper
Lean solution Reflex
filter accumulator
Makeup
Sulfur water
Reflex
absorber
H2S pump
Feed Export
concentrator
gas water
Stripper
Packinox reboiler
exchanger
Figure 10.6 Physical solvent process flow scheme 3: Thermal and Flash Regeneration.
(H2 S) feeds. Because heat energy is not used, flash regen is a low-energy approach
to physical absorption.
Thermal regeneration has been applied to sulfur removal (not CO2 ) and allows a
low (<1 ppm) sulfur specification to be realized. This method works best with high
thermal stability solvents like DMEPG; it is not used with propylene carbonate and
that solvent would degrade under high-temperature regeneration. The use of ther-
mal regen does, of course, add energy cost to the physical absorption process, so the
performance needs of a process need to be balanced with cost considerations.
The third physical absorption scheme includes both thermal and flash distilla-
tion and can accomplish sulfur and CO2 removal. While more complex, a primary
advantage of this approach is that the product acid gas levels can be adjusted to meet
specific customer requirements. This flow scheme is shown in Figure 10.6 and is rep-
resentative of the process design for Selexol physical absorption unit, which will now
be described in more detail.
Methane Hydrogen
H2 S 135 680
CH3 SH 340 1700
CO2 9 45
CO 0.43 2.2
COS 35 175
The selectivity profile of Selexol is summarized in Table 10.9. The high selectivity
of DMEPG of H2 S over CO2 allows efficiency recovery of CO2 in the flash drum (and
minimizes CO2 in acid gas) and saves energy and money in the solvent recovery
process. Likewise, the selectivity of Selexol for CO2 over H2 minimizes downstream
H2 purification requirements.
Table 10.10 summarizes a few of the key advantages of Selexol versus other phys-
ical absorption methods. While most of these advantages speak to lower operating
costs (e.g. less solvent loss, higher availability, and low energy/power requirements),
other advantages such as lower toxicity, lower corrosion (no formation of heat-stable
salts), minimal process effluent, and protection of downstream equipment through
the capture of metal carbonyls add to the benefits of Selexol versus other absorption
options.
Shown in Figure 10.6 is a standard flow scheme for a Selexol unit tasked with the
removal of H2 S, COS, CO2 , HCN, and metal carbonyls. Given the low corrosivity of
DMEPG, a Selexol unit is made mostly of carbon steel.
The first column (sulfur absorber) removes most of the H2 S (and a limited
amount of CO2 ) from the feed syngas, which then flows to the second column
(CO2 absorber), which removes most of the CO2 . The rich solvent leaving the
CO2 absorption column is flashed in drums, from which relatively pure CO2 is
recovered. The solvent in the sulfur absorber column must be stripped in a column
with a reboiler to remove the high H2 S content gases.
10.5 Additional Commercial Uses of Absorption 257
Natural gas
Bulk CO2 removal 12 28–65 0–0.01
Selective H2 S removal 12 2–90 0.7–1.0
Landfill gases
Bulk CO2 removal 9 ∼50 —
Table 10.11 provides some perspective on the number and variety of Selexol units
in operation. As can be seen, Selexol’s combination of selectivity, energy efficiency,
and low corrosivity allows it to be used across a range of gas treatment end uses.
Water vapor
Dry gas out + stripping gas
Flash gas
Still
Glycol
contactor
Flash Reboiler
tank
Stripping
Separator
gas
Wet Surge drum
gas in
Water
Filter
Pump
Figure 10.7 Glycol dehydration unit flow scheme. Source: Bentley [17]. © 1991, American
Institute of Chemical Engineers.
of most of its water content, then is transported out of the top of the contactor
dehydrator.
The glycol solution now contains the water removed from the natural gas and is
processed in a reboiler designed to remove only the water from the solution. The
boiling-point differential between water (100 ∘ C) and glycol (204 ∘ C) makes it rel-
atively easy to remove water from the glycol solution, allowing it be reused in the
dehydration process.
In addition to water from the wet natural gas stream, the glycol solution may
carry with it small amounts of methane and other compounds found in the wet
gas. Previously, this methane was simply vented out of the boiler. In addition to
losing a portion of the natural gas that was extracted, this venting contributes to
greenhouse gas emissions. To decrease the amount of methane and other com-
pounds that are lost, flash tank separator-condensers are employed to remove these
compounds before the glycol solution reaches the reboiler. The flash tank separator
reduces the pressure of the glycol solution stream, which allows the methane and
other hydrocarbons to vaporize (flash). The glycol solution then travels to the
boiler, which may also be fitted with air or water-cooled condensers to capture
any remaining organic compounds in the glycol solution. In total, absorption
systems are capable of recovering 90–99% of the methane that would otherwise
be flared into the atmosphere. The regeneration (stripping) of the glycol is limited
by temperature as both DEG and TEG decompose at or below their respective
boiling points.
References 259
Air Water
References
1 Abu-Zahra, M.R.M., Sodiq, A., and Feron, P.H.M. (2016). Commercial liquid
absorbent based PCC processes. Absorption-Based Post-combustion Capture of
Carbon Dioxide (ed. Paul H.M. Feron), 757–778. Cambridge: Woodhead Publish-
ing.
2 US Energy Information Administration (2020). How much carbon dioxide is pro-
duced when different fuels are burned?, https://www.eia.gov/tools/faqs/faq.php?
id=73&t=11 ()
3 US Energy Information Administration (2019). Natural gas. Monthly Energy
Review (ed. Michael Kopalek), Table 4.3, November 2019, https://www.eia.gov/
totalenergy/data/monthly/.
4 Breeze, P. (2016). The Natural Gas Resource, Gas-Turbine Power Generation, 9–19.
Cambridge, MA: Academic Press.
5 Hammer, G., Lübcke, T., Kettner, R. et al. (2006). Natural Gas. In: Ullmann’s
Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH.
260 10 Absorption
11
Extraction Technology
11.1 Introduction
Though there are many extraction processes, this chapter will focus on liquid–liquid
extraction (LLE) and supercritical fluid extraction (SCFE). Critical factors that con-
trol and affect extraction performance, including solvent selection and type of extrac-
tors, will be described. Selecting the right solvent and type of extractors are key to
achieving high purity and efficiency in the extraction process. After introducing the
fundamentals of extraction, the commercial extraction examples of hydrometallurgy
and hydrocarbon will be introduced.
Product A
Extract A
(contain some
solvent)
Extract
separator
Feed (A+B)
Extractor
Solvent
Raffinate B
Raffinate B separator
(contain some
solvent)
bulk feed. In addition, the solvent phase must be substantially immiscible from
the feed phase. Extraction efficiency can be enhanced by maximizing dispersion
of solvent phase into the feed phase. Extraction columns normally contain trays,
packing, or mechanical agitators to enhance the solvent dispersion. The function
of trays or packing is also to enhance solvent drop formation in the disperse
phase. Decreasing size of solvent drop increases mass transfer efficiency of feed
components to dissolve into the solvent phase. Mass transfer fundamentals in
extraction are like those in distillation.
11.2 Extraction Processes 263
Water
Separator
CO2
Extractor
Caffeine
Coffee Scrubber
beam
CO2
CO2 and
coffeine
CO2 recovery
Figure 11.2 Supercritical extraction of caffeine from coffee beans using carbon dioxide
solvent.
264 11 Extraction Technology
LLE SCFE
11.3 Solvents
The most important component affecting extraction efficiency is the solvent. There-
fore, solvent selection is a critical factor in developing an effective extraction system.
Criteria in solvent selection are described in this section.
11.3.1 Selectivity
Selectivity is calculated from the ratio of the distribution coefficients at equilibrium
of components A and B to be separated. The distribution coefficient of component A
is the ratio of concentration A in the extract phase to concentration A in the raffinate
phase. The distribution coefficient provides a measure of the affinity of component
A in the two phases. For the two-component separation, A and B, the separation
factor or the selectivity of A to B can be calculated as:
Ka
Separation factor or selectivity of A to B =
Kb
Xa
Ka = = distribution coefficient of A at equilibrium
Ya
11.3 Solvents 265
Xb
Kb = = distribution coefficient of B at equilibrium
Yb
Xa = amount of component A in extract phase
11.3.2 Recoverability
After the extraction, the ability to recover solvent from the product is critical to the
efficiency of the process. Typically, this can be done by distillation. Therefore, when
selecting the solvent, the boiling point and thermal stability of the solvent need to
be taken into consideration when designing the system.
11.3.3 Immiscible
An extraction unit, a liquid stream (feed) containing the component(s) to be recov-
ered, is fed into an extractor, where it contacts a solvent. The solvent must be immis-
cible with feed liquid stream. This allows solvent to be dispersed as droplets in the
continued feed liquid phase.
11.3.4 Density
Mass transfer occurs between the droplets (solvent dispersed phase) and the
surrounding liquid (continuous-feed phase). For the two liquids to be subsequently
separated, solvent selected must have different density from the feed liquid stream.
The droplets then accumulate above or below the continuous phase, depending on
the liquids’ relative densities.
11.4 Extractors
There are two types of liquid–liquid extractors: simple extractors and mechan-
ical extractors. Examples of simple extractors include spray, sieve trays, and
packed column. These extractors do not have any moving components. On the
other hand, mechanical extractors, such as mixer-settlers, rotary-disc columns,
reciprocating-plate columns (Karr), and centrifugal extractors, achieve separation
by mechanical agitation. Table 11.2 summarizes the several types of liquid–liquid
extractors.
Extractor Characteristics
Simple extractors
● Spray, sieve, packed column Simple dispersion mechanism; suitable for
only a few stages of separation; low capital cost
Mechanical extractors
● reciprocating-plate column (Karr) Used with high viscosity liquids; system
becomes complex and costly for higher stage of
separation
● centrifugal extractor Used with liquids containing suspended solids;
easy-to-disperse components
Used with emulsified materials and systems
with low-density component differences; short
contact time for unstable materials; high
capital cost
11.4 Extractors 267
Interface
Heavy
Heavy Heavy
phase
phase phase
Packing Packed
Holddown Sections
grid Coalesced
dispersed
Light phase
phase Light Packed Light
phase redistributor phase
Extract
Feed - heavy
(continuous)
Continuous phase
(feed)
Tray
spacing
Solvent - light
(continuous)
Raffinte
In the extraction mechanism described above, the smaller the solvent droplets
generated, the higher the surface area of the dispersed solvent phase. The
smaller droplets result in better mass transfer between the feed and solvent. Mass
transfer efficiency of sieve trays can be predicted using published models and
correlations [1].
Light phase
Heavy phase
Centrifugal extractor
dispersed-phase holdup. Packing also reduces back mixing of the continuous phase
by promoting the breakup of dispersed-phase droplets.
(1) Hydrometallurgy such as extraction of Zr from Hf, and U from Pu; and
(2) Hydrocarbon extraction such as aromatics from nonaromatic, m-xylene from
other xylene isomers, mercaptan removal, and SCFE of caffeine from coffee.
Reaction
Types of extraction solvents mechanism
Scrub:
NaNo3 Product
HNO3 ZrNO3
Feed:
HfNO3 Strip:
ZrNO3 H2O
NaNo3
HNO3
HNO3
Solvent:
TBP
hydrocarbon
Raffinate: (dodecane)
HfNO3 HNO3
NaNo3
HNO3
Key:
organic phase
aqueous phase
and sodium nitrate. The feed is separated into a raffinate stream containing almost
all the hafnium and an extract stream that contains almost all the zirconium. In the
extraction unit, most of the zirconium is extracted into the solvent by complexation
with the TBP solvent. The hafnium is then back extracted from the solvent into the
aqueous phase in the scrubbing section. The scrubbing section is analogous to the
rectification section in a distillation column, and the scrub solution is analogous to
the distillation column’s reflux. The zirconium solute is removed from the solvent in
the stripping section by using a water strip solution. Finally, the solvent is purified
in the solvent regeneration section and then returned to the extraction unit.
Aqueous
Reductant
Pu product
Raffinate: Solvent:
Pu+1,Pu+2,Pu+3 TBP diluent
Herman Pines [5] reports that the basicity of aromatics increases as the number
of alkyl groups bound to the benzene ring increases. Pines’ finding is summarized
in Table 11.5, which shows the relative basicity of methylbenzenes in HF–BF3 sol-
vent. It is clear all the xylenes are more basic than toluene. Furthermore, among the
xylenes, m-xylene is the most basic.
Based on the basicity of the different xylene isomers, Japan Gas–Chemical devel-
oped a LLE process using HF–BF3 solvent to extract m-xylene from another xylene
isomer [6, 7]. In this process, m-xylene (MX) reacts reversibly to form complexes with
boron trifluorides (BF3 ) in the presence of liquid hydrogen fluoride (HF). MX com-
plex (MXHBF4 ) is soluble in HF solvent and then can be separated from the other
feed stream isomers. The MX complex formation with HF–BF3 can be expressed by
274 11 Extraction Technology
Relative
Methylbenzene basicity in HF–BF3
Toluene 0.63
p-Xylene 1
o-Xylene 2
m-Xylene 20
Pseudocumene (1,2,4-) 40
Hemimellitene (1,2,3-) 40
Durene (1,2,4,5-) 120
Prehnitene (1,2,3,4-) 170
Mesitylene (1,3,5-) 2 800
Isodurene (1,2,3,5-) 6 500
Pentamethylbenzene 8 700
Hexamethylbenzene 89 000
(1) With hydrocarbons containing the same number of carbon atoms, the solubility
decreases in the following order: aromatics > naphthenes > olefins > paraffins.
(2) Within the same hydrocarbon homologous series, the solubility decreases as
molecular weight increases.
Due to the unique sulfolane selectivity and solubility properties for hydrocarbons,
the Sulfolane process combines both LLE and extractive distillation (ED) in the same
process unit. This combination of technologies results in the effective processing
of feed stocks with a much broader boiling range than would be possible by either
extraction process alone. This combined extraction process has advantages for aro-
matic recovery in:
(1) LLE systems where light nonaromatic components are more soluble in the sol-
vent than heavy nonaromatics. Thus, LLE is more effective in separating aro-
matics from the heavy contaminants than the light ones.
(2) ED where light nonaromatic components are more readily stripped from the
solvent than heavy nonaromatics. Thus, ED is more effective in separating aro-
matics from light contaminants than heavy ones.
Raffiante
Sulfonate Benzene
extractor
Toluene
H2 LE
C9 aromatic
H2 LE H2 LE
THDA
Naphatha CCR RS H2 LE
HDT platformer Clay BC TC HA
Tatory
C10+
ortho-xylene
Light ends
H2 LE
Clay
Light paraffins +
water
column
Sulfolane +aromatics +
Extract
light paraffins
Oil feed
(Aromatics
Sulfolane +aromatics
+ paraffins)
Water
Sulfolane
Recycle sulfolne
aromatics and nonaromatic components, which are separated using the sulfolane
extraction process.
1 1 1
Caustic regeneration∶ NaRS + O2 + H2 O → NaOH + RSSR
4 2 2
1
Overall reaction∶ 2RSH + O2 → RSSR + H2 O
2
278 11 Extraction Technology
K q – Overall
Mercaptans extraction constant
Ethyl-mercaptan 77.9
n-Propyl-mercaptan 15.6
n-Butyl-mercaptan 3.59, 3.56
Tert-Butyl-mercaptan 2.46
n-Amyl-mercaptan 0.753
Tert-Amyl-mercaptan 0.362
n-Heptyl-mercaptan 0.0342
The extraction efficiency of the Merox process can be characterized by the extrac-
tion coefficient (K q ) in the equilibrium reaction and extraction equation below:
LSR
naphtha FCC Kerosene Distillate
Gas up to 350 °C
LPG Pentane & Naphtha & Jet
phase EP
Condensate
Extractor
Treated
product
Caustic
prewash
Oxidizer Disulfide
separator Spent
air
Disulfide
Air oil
Feed
Caustic settler
Regenerated caustic
Mercaptide-rich caustic
LP stream
Reactor
Dilute Air
caustic
Drain pot
Naphtha Product
To sewer
is for treating naphtha. The Minalk process consists of a single reactor filled with a
fixed bed of activated charcoal impregnated with a Merox proprietary catalyst. Dilute
caustic is continuously injected into the naphtha feed stream along with air. The
naphtha is then passed downward through the fixed bed where the mercaptan is
oxidized to disulfide. The disulfide remains in the naphtha-phase product.
11.6 Summary
LLE, or solvent extraction, is one of the oldest and widely used techniques in indus-
trial separation. Supercritical extraction was developed in the later days with more
uses being developed. This chapter summarized the key components of the process,
including how to choose solvents and type of extractors. In addition to the basic prin-
ciples of extraction, the chapter highlighted some of the key commercial extraction
processes in metal extraction and hydrocarbon separation.
Acknowledgments
The author would like to thank Dr. Ames Kulprathipanja for reviewing the chapter.
Thanks also to Dr. Katipot Inkong and Ms. Orawee Lamoonkit of Chulalongkorn
University in drawing some of the figures.
References 281
References
1 Rocha, J.A., Humphrey, J.L., and Fair, J.R. (1986). Mass transfer efficiency of
sieve tray extractors. Industrial and Engineering Chemistry Process Design and
Development 25: 286.
2 Ritcey, G.M. and Ashbrook, A.W. (1984). Solvent Extraction, Principles and Appli-
cations to Process Metallurgy. New York: Elsevier.
3 Cox, R.P., Peterson, H.C., and Beyer, G.H. (1958). Separating hafnium from zir-
conium. Solvent extraction with tributyl phosphate. Industrial and Engineering
Chemistry 50 (2).
4 Schultz, W.W. and Navratil, J.D. (1984). Science and Technology of Tributyl Phos-
phate. Boca Raton, FL: CRC Press.
5 Pines, H. (1981). The Chemistry of Catalytic Hydrocarbon Conversions, 26.
Academic Press.
6 Ueno, T. (1970). Japan Gas-Chemical Xylene Separation Process, vol. 12. Bulletin
of The Japan Petroleum Institute.
7 Talbot, J.L. (1956). Separation of xylene isomers. US Patent 2,738,372.
8 Neuzil, R.W. (1982). US Patent 4,326,092.
9 Kulprathipanja, S. (1999). Process for adsorptive separation of metaxylene from
xylene mixtures. US Patent 5,900,523.
10 Kelly, M.F., and Uitti, K.D. (1970). Extractive distillation of aromatics with
sulfolane solvent. US Patent 3,551,327.
11 Meyers, R.A. (2016). Handbook of Petroleum Refining Processes, Fourthe. UOP
Sulfolane Process: McGraw-Hill Professional.
12 Yabroff, D.L. Extraction of mercaptans with alkaline solution. Industrial and
Engineering Chemistry 32 (2).
13 Sullivan D., The Role of the Merox Process in the Era of Ultra Low Sulfur Trans-
portation Fuels. Fifth EMEA Catalyst Technology Conference, 3 and 4 March
2004
283
Index
a zeolite 148
absorption zeolite powders 152–153
acid gas removal 243–245 zeolite synthesis 150–152
air stripping 259 adsorption
chemical 241 adsorbent 148–150
chemical solvent 245–251 bulk liquid adsorptive separation
CO2 emissions 241 154–170
gas separation 243 commercial bulk liquid adsorptive
glycol dehydration 257–258 process 170–174
natural gas 241, 242 principle of 147–148
physical 241 adsorption isotherms 156
physical solvent 251–257 adsorption step 154
accretion of zeolite powder 153–154 air stripping 259
acid gas removal 46, 63, 65, 66, 67, 228, alkanes 4, 5, 80, 197
243–245, 247, 257 alkyd resins 133, 139
acrylamide 81 alkylation 7, 8, 27, 29, 33, 34, 56, 58, 59,
acrylic acid 25, 75 103, 122, 128, 197, 226
acrylic resins 129, 139 alkylbenzene sulfonates 132–133
acrylonitrile 25, 30, 45, 81, 129, 198, 234 alkylbenzenes 132–133
Acrylonitrile–Butadiene–Styrene (ABS) AlPO-14/polymer MMMs 220
30, 81, 129 α-alumina porous support membrane
activated alumina 148 218
activated carbon 148, 219 alumina-supported platinum catalysts
activated clay 148 51
activity coefficients 185 aluminosilicates 60, 148, 153, 169, 214
adipic acid 34, 124–127 Amine Guard Formulated Solvent (AGFS)
adipic acid production 124–128 249
adiponitrile 81 Amine GuardTM FS process 63
adsorbate 147, 148, 154, 156, 157, 159, amines process 246–248
160, 162, 164, 166, 168 ammonia 13, 37, 38–39, 45, 67, 75, 81,
adsorbents 82–83, 89, 114, 180, 204, 228, 249,
general uses of 148–149 250
pore size distribution 148 aniline 113, 130, 131, 180
iso-paraffins 47, 167, 197 liquid hydrogen fluoride (HF) 59, 162,
isophthalic acid 133, 134, 174 174, 272–274
isopropylbenzene 128, 166–168 liquid petroleum 4–5
isotactic polypropylene 79 liquid-liquid extraction (LLE) 261–262,
I-400 catalyst 102 264, 275, 278–279
vs. supercritical fluid extraction (SCFE)
j 264
jet fuel 3, 10, 19, 59, 197, 277, 278 Li-zeolite 161
Joule-Thompson effect 112 low pressure polyethylene 75
low-density polyethylene (LDPE) 75, 76
k LPG 7, 10, 19, 52, 88, 105, 107, 109, 111,
KataMaxTM structural packing system 119, 197, 277, 278
200 lubricants 3, 153
kerosene 3, 7, 10, 45, 57, 58, 59, 167, lubricating oils 10, 264
197, 277, 278
K-zeolite 161 m
maleic anhydride 34, 133, 138
l m-aminophenol 130
LAB technology 59 MaxEne process 47, 48
light crude oil 4–5 Maxwell model 220
Light DesorbentParex 172–174 McCabe-Thiele design method 182
light distillates 8, 10 McCabe-Thiele diagram 182–184
light gas oil 3 m-chloroaniline 130
light naphtha 19, 88, 105–107, 109, 277, MDI 130–132
278 mechanical extractors 266, 268
light olefins 45, 48, 49, 62 medium conversion refinery 8, 9
light paraffin 45, 48–57 medium-pore zeolite 60
light paraffin dehydrogenation membrane contactor 204, 207, 227, 234,
iso-C4 Oleflex complex 54–56 235
process and catalyst 48–51 membrane permeance 206, 217, 218
UOP C3 Oleflex Complex 51–54 membrane separation 203, 206, 211,
UOP C3 /C4 Oleflex complex 56 212, 216, 223, 225, 227, 228, 229,
UOP mixed-C4 Oleflex complex 56–57 233, 234
light solvent 261 membranes 64
light/heavy naphtha 3 gas separation 228
linear alkylbenzene (LAB) 45, 57–59, inorganic 213–218
75, 197 ion-exchange 235
linear low-density polyethylene (LLDPE) membrane contactor 234–235
75–77 microfiltration 234
linear-alkylbenzene 132 mixed matrix 218–223
liquefied natural gas (LNG) 63 modules 223–225
liquefied petroleum gas 3, 88, 277 nanofiltration 233–234
liquid adsorption 148, 154, 160 pervaporation 233
Liquid Hourly Space Velocity (LHSV) polymeric 205–207
93 principle and background 203–204
290 Index
styrene-butadiene (SB) copolymer 30, tertiary amyl methyl ether (TAME) 27,
122 40, 54
styrene-butadiene rubber (SBR) 30 tetrahydrothiophen 6, 275
styrenics 21–22, 33–34 tetrahydrothiophene-1,1-dioxide 275
sulfolane process 110, 272, 274–275, 277 tetramethylbenzene 103
SulfolaneTM extractive distillation unit thermal dealkylation process (THDA)
172 95
sulfur compounds 6, 72 thermal regeneration 251, 254–255
sulfur-containing molecules 10 thermally-derived ethylene 19
superabsorbent polymers 25 thermocracking 8
supercritical fluid extraction (SCFE) thermodynamic vapor-liquid equilibrium
261, 263–264 (VLE) compositions 181
liquid-liquid extraction (LLE) 264 thin film composite (TFC) membranes
surface area 51, 89, 91–92, 148, 203, 207
153–154, 184–187, 190, 195, 206, flat sheet membrane 207–208, 227
hollow fiber membrane 207–208
215–217, 234, 243, 267
polymeric membranes 211–213
sweetening Merox process 278–280
structure 207
symmetric nonporous dense film
titanium silicates 214
polymeric membranes 207
toluene 3, 11, 17, 32–36, 87, 89, 95–105,
symmetric porous polymeric membranes
111–115, 118, 121, 127, 134, 162,
207
171–174, 197–199, 257, 264, 273
syndiotactic polypropylene 79
toluene diisocyanate (TDI) 35, 131–132
syngas 75
toluene methylation 96, 103–105
ammonia 82–83
toluene methylation unit 104
hydrogen 82
toluene xylenes 6, 197
methanol 83 toluol 11
syngas production topping refineries 8
coal gasification 67–72 transalkylation 32–33, 36–37, 95–100,
natural gas conversion 64–67 173
removal of impurities in raw natural gas Tray hydraulics 186
63–64 tray vs. packed columns 187, 188
synthesis gas 3, 11, 17 tri-ethylene glycol 78
ammonia 38–39 2,2,4-trimethylpentane 5
methanol 39–40 3,3,5-trimethylheptane 167
synthesis gas complexes 13 tri-substituted paraffin 167
synthetic fibers 133, 142, 170 tributyl phosphate (TBP) 270–273
synthetic gas (syngas) 45 solvent 272–273
triethanolamine (TEA) 151, 247
t trimethylbenzene 103
tacticity 79 trinitrotoluene 87
TatorayTM transalkylation process 96 typical naphtha-based olefins unit 12
terephthalic acid 12, 37, 133, 139
tertiary amines 244, 246, 248–249 u
tertiary amyl ethyl ether (TAEE) 54 Udex 110
296 Index