Chapter 3

Industrial Organic Chemistry

Abstract

The chemical process industries play an important role in the development of a country by providing a wide variety of products and use raw material derived from petroleum and natural gas, salt, oil and fats, biomass and energy from coal, natural gas and a small percentage from renewable energy resources. Although manufacture of organic chemicals initially started with coal and alcohol from fermentation industry, later due to availability of petroleum and natural gas which dominated the scene, now more than 90% of organic chemicals are produced from petroleum and natural gas routes. However, rising cost of petroleum and natural gas and continuous decrease in the reserves has spurred the chemical industry for alternative feedstocks like coal, biomass, coal bed methane, shale gas, and sand oil as an alternate source of fuel and chemical feedstock.

This chapter deals with the foundations of organic chemistry from an industrial perspective and presents the various processes that are produced and on a regular (almost a day-to-day) basis. This will give the reader the ability to understand the necessary links between laboratory organic chemistry and industrial process chemistry that is a necessary and growing phenomenon within the chemistry community. These chemistry and engineering sectors of industry have long held strong ties since chemistry points the way to synthetic pathways and engineering points that way by which these pathways might be achieved on a commercial scale.

Keywords

Organic chemicals; Crude oil; Coal; Natural gas; Biomass; C-1 chemistry; C-2 chemistry; C-3 chemistry; C-4 chemistry; BTX chemistry

1 Introduction

The chemical process industries play an important role in the development of a country by providing a wide variety of products and use raw material derived from petroleum and natural gas, salt, oil and fats, biomass and energy from coal, natural gas and a small percentage from renewable energy resources. Although manufacture of organic chemicals initially started with coal and alcohol from fermentation industry, later due to availability of petroleum and natural gas dominated the scene and now more than 90% of organic chemicals are produced from petroleum and natural gas routes. However, rising cost of petroleum and natural gas and continuous decrease in the reserves has spurred the chemical industry for alternative feedstocks like coal, biomass, coal bed methane, shale gas, and sand oil as an alternate source of fuel and chemical feedstock.

However, scientific and engineering interests have not always been aligned and, therefore, maintaining a partnership between chemists and engineers with the focus on industrial processes and the production of organic chemicals is an opportunity for both disciplines that and can create a true symbiosis. This type of collaboration is engendered by the desire of industrial operations to access the chemical and engineering expertise that is required for the development of new technologies at the forefront of process innovations. From both, the chemical standpoint and the engineering standpoint, the symbiotic relationship represents the perfect occasion to apply scientific and engineering research concepts to solve problems that benefit a variety of processes.

Furthermore, the umbrella term the industrial organic chemicals industry includes thousands of chemicals and hundreds of processes. In general, a set of building blocks (feedstocks) is combined in a series of reaction steps to produce both intermediates and end-products. Moreover, organic chemicals, particularly petrochemicals, play an indispensable role in the modern world. They are essential ingredients in plastics, synthetic fibers, rubber, fertilizers, and chemical intermediates, which are converted into a wide range of industrial products. They are the primary building blocks of important materials supporting the health, food, transportation, and communication industries. Organic substances also have made possible many important specialty items, such as protective clothing and materials used for space exploration. The primary organic chemical building blocks (generated principally from petroleum refining)—using benzene, toluene, and the xylene isomers as examples (Fig. 3.1)—are a key subset of the large volume secondary building blocks and a set of large volume tertiary building blocks that participate in a variety of reaction types and a large variety of processes that are used in manufacture of organic chemicals.

f03-01-9780128044926
Fig. 3.1 Selected examples (benzene, toluene, and the xylenes) of the organic chemical building blocks.

The industrial organic chemical sector produces organic chemicals (chemicals that contain carbon) used as either chemical intermediates or end-products. The industrial organic chemical industry uses feedstocks derived from petroleum and natural gas (about 90%) and from recovered coal tar condensates generated by coke production (about 10%). The chemical industry produces raw materials and intermediates, as well as a wide variety of finished products for industry, business, and individual consumers. Important classes of chemicals produced by organic chemical industry facilities include: (1) noncyclic organic chemicals such as acetic, chloroacetic, adipic, formic, oxalic acids and their metallic salts, chloral, formaldehyde, and methylamine; (2) solvents such as amyl, butyl and ethyl alcohols; methanol; amyl, butyl, and ethyl acetates; ethyl ether, ethylene glycol ether and diethylene glycol ether; acetone, carbon disulfide, and chlorinated solvents such as carbon tetrachloride, tetrachloroethylene, and trichloroethylene; (3) polyhydric alcohols such as ethylene glycol, sorbitol, pentaerythritol, and synthetic glycerin; (4) synthetic perfumes and flavoring materials such as coumarin, methyl salicylate, saccharin, citral, citronellal, synthetic geraniol, ionone, terpineol, and synthetic vanillin; (5) rubber processing chemicals such as accelerators and antioxidants, both cyclic and acyclic; (6) plasticizers, both cyclic and acyclic, such as esters of phosphoric acid, phthalic anhydride, adipic acid, lauric acid, oleic acid, sebacic acid, and stearic acid; (7) synthetic tanning agents such as sulfonic acid condensates; and (8) esters and amines of polyhydric alcohols and fatty and other acids.

Relatively few organic chemical manufacturing facilities are single-product (or single-process) plants and many process units are designed to be flexible (with options) so that production levels of related products (or by-products) can be varied over wide ranges. This flexibility is required to accommodate variations in feedstock purity properties which can change the production rate and processes used, even on a short-term (less than a year) basis. Furthermore, the process by-products are also valuable saleable products and the value of these by-products can change the economics of the process.

The typical chemical synthesis process involves combining multiple feedstocks in a series of unit operations—the petroleum refining industry is the best example of the integration of a series of unit processes to produce the desired products (Speight, 2014a, 2016). This is not always the case in the organic chemicals industry where the first unit operation is typically a chemical reaction to produce the product or an intermediate product that leads to the desired chemical. In addition, there is some differentiation between the chemicals to be produced—commodity chemicals tend to be synthesized in a continuous reactor while specialty chemicals usually are typically produced in a batch reactor. Many, but not all, reactions (1) take place at high temperatures, (2) involve metal catalysts, and (3) include one or two additional reaction components. The yield of the organic chemical will/may determine the quantity, type and by-products, including gaseous emissions. The production of many organic specialty chemicals often requires a series of two or more reaction steps and once the reaction is complete, the desired product must be separated from the by-products by a second unit operation. In the separation stage, a number of separation techniques such as settling, distillation or refrigeration may be used and the final product may require further processing (for example, by spray drying or pelletizing) to produce the saleable item. The separation technology employed depends on many factors including the phases of the substances being separated, the number of components in the mixture, and whether recovery of by-products is important. Numerous techniques such as distillation, extraction, filtration, and settling can be used singly or in combination to accomplish the separation.

Finally, regulatory laws regarding hazardous emissions generated during the production of organic chemicals are important dynamics that shape the industry. To minimize the detrimental effects of chemical industry pollutants, multiple local, state, and federal laws govern producers. For example, the federal Emergency Planning and Community Right-to-Know Act requires many manufacturers to submit details of any emissions data to the United States Environmental Protection Agency (US EPA). Similarly, the Pollution Prevention Act requires those same companies to report their waste management and pollution reduction activities. Other federal regulations impacting producers include the Safe Drinking Water Act, the Clean Air Act and Amendments, and other laws that restrict hazardous wastes. In addition to legal restrictions, both the US EPA and the Chemical Manufacturers Association (CMA) sponsor successful voluntary pollution reduction programs that encourage environmental sensitivity. The US EPA has continued to monitor the industry and in the light of current strong emphasis on chemical safety and pollution controls, it is likely that regulations will continue to be added and modified.

There are also voluntary programs where the member companies work with the public to address such issues as chemical safety. The mechanism for these programs involves a combination of soliciting information from the public about the various concerns which are addressed and the progress is reported back to the public. While increasing federal and state regulations pose an ongoing challenge to chemical industry participants, positive signs have indicated that the organic chemicals industry has been successful in clearing these hurdles. Overall chemical industry has managed to reduce the various emissions of waste that appear on the Toxics Release Inventory.

The previous chapter (Chapter 2) has introduced the reader to the varied, but fundamental, aspects of organic chemistry. However, organic chemistry as practiced on the industrial stage is not so simple and/or straightforward. Industrial organic chemistry is an extremely comprehensive and practical discipline and, although work there benefits from understanding the basic organic chemical science, there is still the need to gain a valuable insight into chemical technology. Basic organic chemistry does provide but other chemicals used and produced in industrial processes offer a considerable (but valuable) challenge to (1) an understanding of the processes, (2) the process parameters, (3) the properties of the feedstocks, (4) the properties of the products, (5) the properties of the by-products, and (6) the influence of these various chemicals on the environment. These effects are difficult to understand on the basis of laboratory chemical studies alone and it is for this reason that this chapter is included in the book.

Thus, this chapter deals with the foundations of organic chemistry from an industrial perspective and presents the various processes that are used on a regular (almost a day-to-day) basis. This will give the reader the ability to understand the necessary links between laboratory organic chemistry and industrial process chemistry that is a necessary and growing phenomenon within the chemistry community. These chemistry and engineering sectors of industry have long held strong ties since chemistry points the way to synthetic pathways and engineering points that way by which these pathways might be achieved on a commercial scale.

2 Production of Organic Chemicals

The typical organic chemical synthesis process involves combining one or more feedstocks in a series of unit process operations. Commodity chemicals tend to be synthesized in a continuous reactor while specialty chemicals usually are produced in batches. Most reactions take place at high temperatures, involve metal catalysts, and include one or two additional reaction components. The yield of the organic chemical will partially determine the kind and quantity of by-products and releases. In fact, many specialty organic chemicals require a series of two or three reaction steps, each involving a different reactor system and each capable of producing by-products. Once the reaction is complete, the desired product must be separated from the by-products by a second unit operation. A number of separation techniques such as settling, distillation or refrigeration may be used. The final product may be further processed, such as by spray drying or pelletizing, to produce the saleable item. Frequently, by-products are also sold and their value can influence the economics of the process.

Feedstock costs are the highest variable cost in production of the organic commodity chemicals. The larger producers integrate feedstocks and derivatives production in order to minimize production costs and price fluctuations. Smaller firms do not possess this integration flexibility, making them more susceptible to variations in feedstock price swings. When feedstock prices rise, manufacturers often lower operating rates or suspend production if price increases are not possible. Some producers have the ability to switch feedstocks in order to obtain better feedstocks which lead to better market prices. Moreover, the type of reaction process used to manufacture chemicals depends on the intended product.

Because of the heightened interest in alternative fuels production that has expanded over the past three decades, the use of alternative feedstocks for production of the high-volume commodity chemicals has been, and continues to be, evaluated. Petroleum is still a major feedstock for the production of organic chemicals. However, the currently high prices for petroleum and natural gas have spurred the US chemical industry to evaluate alternative feedstocks for the production of commodity chemicals. These feedstocks include unconventional processing technologies, such as (1) the increased use of natural gas, (2) coal, which includes: coal gasification and coal liquefaction, (3) heavy oil, (4) bitumen from tar sand formations, (5) liquids from oil shale processing, and (6) novel resources such as the various types of biomass. Thus, as part of this increased interest, new pathways for commodity chemicals manufacture are continually being developed using feedstocks that offer alternatives to petroleum as well as the development of the varying process options from production of the products.

Finally, there is the need to realize that crude oil has become integral to the necessities of the modern world. However, although the era of crude oil may be drawing to a close within the next 50 years or so (Hirsch et al., 2006; Speight, 2011a, b; Speight and Islam, 2016), petrochemicals have remained the mainstay of the chemical and energy industries and will continue to do so during the next five decades. The infrastructure of the chemicals industry remains focused on crude oil and natural gas and any alternative must work within that system and there is a much better chance of making the transition to biological sources if new processes are integrated into existing technology. However, despite the efforts to facilitate biomass conversion, there have been tangible successes but renewable feedstocks are not yet fulfilling their potential to replace crude oil and natural gas. In addition, biomass derivatives rarely have the purity and homogeneity that crude oil and natural gas feedstocks offer for production of organic chemicals and, in addition, purification processes for bioderived streams may be particularly difficult to process. The challenge is to transform low cost bioderived feedstocks into high-value derivatives with commercial applications. Currently, crude oil, natural gas liquids including gas condensate, and natural gas account for the majority of the feedstock materials used by the organic chemicals industry. Natural gas is predominately used to manufacture methanol and ammonia, and the majority of the olefin productions (particularly ethylene) production is based on natural gas liquids. Coal, once the prime feedstock for the chemicals industry became a lesser feedstock for chemical production after World War II when crude oil and natural gas, being more easy to convent to organic chemicals, became the leading feedstocks for the chemical industry.

The price volatility of crude oil and natural gas and the possible dearth of these feedstocks in the distant future (Speight, 2011a, b; Speight and Islam, 2016) have spurred the chemical industry to examine alternative feedstocks for the production of commodity chemicals. Thus, over the last 30 years, alternatives to conventional petroleum and natural gas feedstocks have been developed, but have limited, if any, commercial implementation. Alternative feedstocks under consideration include coal from known processing technologies, such as gasification and liquefaction, novel resources such as biomass, heavy oil, and bitumen from tar sands or oil shale.

Sources of organic compounds, such as ethanol from sugar fermentation and bitumen-derived heavy crude oil are now being primarily exploited for fuels, rather than for chemical feedstocks. In fact, over the last 50 years, there has been much activity in the development of alternative feedstocks, but little activity in bringing technologies to market—the delay in implementation of any such technologies rests on the economics of the process. The economic competitiveness of technologies such as integrated gasification combined cycle (IGCC) and gas-to-liquids (GTL) depend on the current and predicted prices of crude oil and alternative feedstocks, and the costs for transportation and storage.

Most of the work into alternative feedstocks has focused on energy production, either for electricity, liquid fuels from synthesis gas (syngas, a mixture of carbon monoxide and hydrogen), or bioethanol (i.e., ethanol from biomass). Some technologies for chemicals production are mature, such as coal gasification, and are ready for implementation if economically feasible. Other unconventional sources of organic compounds, such as ethanol from sugar fermentation, are now being exploited for fuels, rather than for chemical feedstocks.

This section presents a review of the various types of feedstocks that are used for production of organic chemicals (single chemicals and mixtures of chemicals such as petroleum products) and the means by which these feedstocks are used to produce chemical products and how they fit into the current future production of organic chemicals.

2.1 Chemicals From Petroleum

The organic commodity chemicals are a group of crude oil-derivative chemicals (also known as petrochemicals) used as intermediates to produce other chemicals, which, in turn, are used to manufacture a wide variety of end-use products, including construction materials, apparel, adhesives, plastics, and tires. The majority of the organic commodity chemicals are derived from benzene, a chemical derived from crude oil refining. Examples of specific compounds in this hydrocarbon group include ethylbenzene (C6H5C2H5), styrene (C6H5Cglyph_dbndH2), cumene [C6H5CH(CH3)2], ortho-xylene (0-1,2-CH3C6H4CH3), meta-xylene (1,3-CH3C6H4CH3), and para-xylene (1,4-CH3C6H4CH3) while (C6H5OH) and aniline (C6H5NH2) are products that contain the oxygen and nitrogen heteroatoms, respectively.

As commodities, the chemicals produced by one manufacturer should be indistinguishable from those same chemicals produced by another manufacturer, given the same levels of purity. This similarity of products allows consumers to purchase similar product from a wide variety of suppliers, making price the dominant economic factor in purchasing decisions. But it should be noted that some manufactures sell chemical products that may not be 100% pure, and the level of purity must be stated on the packaging.

Chemicals from crude oil usually take the form of products that are mixtures or petrochemicals in which the product is typically an identifiable single organic compound. Petroleum products in contrast to petrochemicals, are those bulk fractions that are derived from petroleum and have commercial value as a bulk product. In the strictest sense, petrochemicals are also petroleum products but they are individual chemicals that are used as the basic building blocks of the chemical industry (Speight, 2011a,b, 2014a).

The constant demand for products, such as liquid fuels, is the main driving force behind the petroleum industry. Other products, such as lubricating oil, wax, and asphalt, have also added to the popularity of petroleum as a national resource. Indeed, fuel products that are derived from petroleum supply more than half of the total supply of energy use on a worldwide basis. Gasoline, kerosene, and diesel oil provide fuel for automobiles, tractors, trucks, aircraft, and ships. Fuel oil is used to heat homes and commercial buildings, as well as to generate electricity. Furthermore, petroleum products are the basic materials used for the manufacture of synthetic fibers for clothing and in plastics, paints, fertilizers, insecticides, soaps, and synthetic rubber. The uses of crude oil as a source of raw material in manufacturing are central to the functioning of modern industry.

Unlike processes, products are more difficult to be placed on an individual evolutionary scale. Processes changed and evolved to accommodate the demand for, say, higher-octane fuels, longer-lasting asphalt, or lower sulfur coke. Another consideration that must be acknowledged is the change in character and composition of the original petroleum feedstock. In the early days of the petroleum industry several products were obtained by distillation and could be used without any further treatment. In the modern refinery, the different character and composition of the petroleum dictates that any liquids obtained by distillation must go through one or more of the several available product improvement processes. Such changes in feedstock character and composition have caused the refining industry to evolve in a direction such that changes in the petroleum can be accommodated.

There is a myriad of products that have evolved through the life of the petroleum industry and the complexities of product composition have matched the evolution of the products. In fact, it is the complexity of product composition that has served the industry well and, at the same time, had an adverse effect on product use. Product complexity has made the industry unique among industries. Product complexity, and the means by which the product is evaluated, has made the industry unique among all industries. But product complexity has also brought to the fore issues such as instability and incompatibility. In order to understand the evolution of the products it is essential to have an understanding of the composition of the various products.

In the simplest sense, naphtha contains varying amounts of paraffins, olefins, naphthene constituents, and aromatics and olefins in different proportions in addition to potential isomers of paraffin that exist in naphtha boiling range. As a result, naphtha is divided predominantly into two main types: (1) aliphatic naphtha and (2) aromatic (naphtha). The two types differ in two ways: first, in the kind of hydrocarbons making up the solvent, and second, in the methods used for their manufacture. Aliphatic solvents are composed of paraffinic hydrocarbons and cycloparaffins (naphthenes), and may be obtained directly from crude petroleum by distillation. The second type of naphtha contains aromatics, usually alkyl-substituted benzene, and is very rarely, if at all, obtained from petroleum as straight-run materials. The products that are higher boiling than naphtha contain higher molecular weight constituents that vary in molecular type.

The high-boiling and more complex lubricating oil is distinguished from other fractions of crude oil by their usually high (> 400°C, > 750°F) boiling point, as well as their high viscosity. Materials suitable for the production of lubricating oils are comprised principally of hydrocarbons containing from 25 to 35 or even 40 carbon atoms per molecule, whereas residual stocks may contain hydrocarbons with 50 or more (up to 80 or so) carbon atoms per molecule. The composition of lubricating oil may be substantially different from the lubricant fraction from which it was derived, since wax (normal paraffins) is removed by distillation or refining by solvent extraction and adsorption preferentially removes nonhydrocarbon constituents as well as polynuclear aromatic compounds and the multiring cycloparaffins (Speight, 2014a, 2016).

There are general indications that the lubricant fraction contains a greater proportion of normal and branched paraffins than the lower boiling portions of petroleum. For the polycycloparaffin derivatives, a good proportion of the rings appear to be in condensed structures, and both cyclopentyl and cyclohexyl nuclei are present. The methylene groups appear principally in unsubstituted chains at least four carbon atoms in length, but the cycloparaffin rings are highly substituted with relatively short side chains. Mono-, di-, and trinuclear aromatic compounds appear to be the main constituents of the aromatic portion, but material with more aromatic nuclei per molecule may also be present. For the binuclear aromatics, most of the material consists of naphthalene types. For the trinuclear aromatics, the phenanthrene type of structure predominates over the anthracene type. There are also indications that the greater part of the aromatic compounds occur as mixed aromatic-cycloparaffin compounds.

After lubricating oil, there is paraffin wax which is a solid crystalline mixture of straight-chain (normal) hydrocarbons ranging from C20 to C30 and possibly higher, that is, CH3(CH2)nCH3 where n ≥ 18. Wax is distinguished by its solid state at ordinary temperatures (25°C, 77°F) and low viscosity when melted. However, in contrast to petroleum wax, petrolatum (petroleum jelly), although solid at ordinary temperatures, does in fact contain both solid and liquid hydrocarbons. It is essentially a low-melting, ductile, microcrystalline wax.

Another product, asphalt which is used as a mastic in various applications (such as road construction and repair), is the residue of mixed-base and asphalt-base crude oils. It cannot be distilled even under the highest vacuum, because the temperatures required to do this promote formation of coke. Asphalt have complex chemical and physical compositions that usually vary with the source of the crude oil.

The final product, coke is the residue left by the destructive distillation of petroleum residua. That formed in catalytic cracking operations is usually nonrecoverable, as it is often employed as fuel for the process. The composition of petroleum coke varies with the source of the crude oil, but in general, large amounts of high-molecular-weight complex hydrocarbons (rich in carbon but correspondingly poor in hydrogen) make up a high proportion. The solubility of petroleum coke in carbon disulfide has been reported to be as high as 50–80%, but this is in fact a misnomer, since the coke is the insoluble, honeycomb material that is the end product of thermal processes.

2.2 Chemicals From Natural Gas

The principal constituent of natural gas is methane (CH4). Other constituents are paraffinic hydrocarbons such as ethane (CH3CH3), propane (CH3CH2CH3), and the butanes [CH3CH2CH2CH3 and/or (CH3)3CH]. Many natural gases contain nitrogen (N2) as well as carbon dioxide (CO2) and hydrogen sulfide (H2S). Trace quantities of argon, hydrogen, and helium may also be present. Generally, the hydrocarbons having a higher molecular weight than methane, carbon dioxide, and hydrogen sulfide are removed from natural gas prior to its use as a fuel. Gases produced in a refinery contain methane, ethane, ethylene, propylene, hydrogen, carbon monoxide, carbon dioxide, and nitrogen, with low concentrations of water vapor, oxygen, and other gases.

Transportation of natural gas from isolated sources could best be effected by converting the light hydrocarbons to easily transportable liquids, such as methanol, in situ. Natural gas is used to manufacture methanol offshore, but production in the United States has been curtailed because of the high cost of natural gas. Methanol is a key building block chemical, and a methanol-based economy has been touted as an alternative to a hydrogen economy. Technologies exist for methanol-to-olefin production that are ready for implementation. These processes will become viable depending on the prices of the feedstocks and oil, all of which are highly variable, as well as costs for transportation and storage.

Unconventional or stranded natural gas is being evaluated as an alternative to alleviate current shortages in natural gas supply. Stranded, or unconventional, natural gas is that which is not easily transported from source to end use by pipeline, or is uneconomical to transport as a gas. Unconventional natural gas may come from a variety of sources: natural gas hydrates, stranded or geographically remote methane, such as that from sites in Alaska or the Rocky Mountains, coal bed methane, and methane from anaerobic fermentation such as occurs in landfills.

Gas from unconventional sources must be controlled, concentrated, or converted to a liquid form or stabilized in some manner, before being transported to the chemical manufacturer, power plant, or end user. The conversion of methane to a liquid product, such as methanol (CH3OH), will allow easier transportation of materials from remote sites.

Liquid products can be made from natural gas through a process involving the conversion of synthesis gas (syngas). Some synthesis routes, such as the production of diesel fuel, are well understood (Chadeesingh, 2011) and one example of gas to liquids is the Fischer-Tropsch production of liquid fuels or methanol to gasoline (MTG). Methanol can give rise to a number of building block organic chemicals such as: acetic acid (CH3CO2H), dimethyl ether (CH3OCH3), formaldehyde (HCHO), ethane (C2H6), and propane (CH3CH2CH3).

2.3 Chemicals From Coal

Substantial worldwide coal reserves make coal an attractive alternative to natural gas and petroleum. Historically, research into coal gasification has been focused on energy fuels, and more recently on power production, with less emphasis on commodity chemicals production. Efforts to develop pathways for chemicals production using optimal catalysts and for process scale-up are needed to replace conventional petroleum. The Fischer-Tropsch synthesis hydrocarbons (Chadeesingh, 2011) is particularly useful as a means of converting coal (via the production of synthesis gas) to commodity chemicals. While this pathway to organic chemical is different to the pathways used for crude oil and crude oil products, they are an option for use of a plentiful resource (coal). In fact, there have been several assessments of the use of coal for the generation of synthesis gas and liquid fuels through Fischer-Tropsch process.

2.3.1 Gasification

Organic chemicals can be produced through the gasification of coal. Because of the large domestic reserves of coal on a worldwide basis, this feedstock option is one of the strongest for chemical production in the long term. The technology associated with the gasification of coal [and other carbonaceous feedstocks is well understood (Chadeesingh, 2011)], having been used in Germany in WWII and in South Africa to produce liquid fuels in combination with Fischer-Tropsch processing.

The typical gasification process starts with the production of synthesis gas (a mixture of carbon monoxide and hydrogen) in the gasifier:

Ccoal+H2OH2+COCO+H2OH2+CO2

si1_e

Water-gas shift reaction:

CO+H2OCO2+H2

si2_e

A generic oxygen gasification system comprises the following main steps: (1) reaction of the coal with oxygen or steam at 1000 + psi, (2) quench with water to remove particles and cool the synthesis gas, (3) application of the water-gas-shift reaction to produce hydrogen and carbon dioxide, and (4) application of gas cleaning technology by the use of physical solvents for the simultaneous removal of hydrogen sulfide and carbon dioxide (Mokhatab et al., 2006; Speight, 2007, 2013, 2014a, b). The solvents are recovered by depressurization and the hydrogen sulfide is converted to sulfur in a two-step process by first heating in oxygen to produce sulfur dioxide which then reacts further with hydrogen sulfide to produce sulfur and steam. Further catalysis increases the production of hydrogen by the reaction of carbon monoxide to carbon dioxide in the presence of water to produce additional hydrogen. The Fischer-Tropsch process can be used to produce alkanes, which are building blocks for many large-volume chemicals.

In addition to the production of synthesis gas, other products of coal gasification which may find use in the organic chemicals industry are (1) producer gas, which is a low Btu gas obtained from a coal gasifier (fixed-bed) upon introduction of air instead of oxygen into the fuel bed—the composition of the producer gas is approximately 28% v/v carbon monoxide, 55% v/v nitrogen, 12% v/v hydrogen, and 5% v/v methane with some carbon dioxide, (2) water-gas, which is a medium Btu gas that is produced by the introduction of steam into the hot fuel bed of the gasifier—the composition of the gas is approximately 50% v/v hydrogen and 40% v/v carbon monoxide with small amounts of nitrogen and carbon dioxide, (3) town gas, which is a medium Btu gas that is produced in the coke ovens and has the approximate composition: 55% v/v hydrogen, 27% v/v methane, 6% v/v carbon monoxide, 10% v/v nitrogen, and 2% v/v carbon dioxide—carbon monoxide can be removed from the gas by catalytic treatment with steam to produce carbon dioxide and hydrogen, and (4) synthetic natural gas (SNG), which is methane obtained from the reaction of carbon monoxide or carbon with hydrogen—depending on the methane concentration, the heating value can be in the range of high-Btu gases.

Because any organic carbonaceous material can be gasified (Speight, 2013, 2014b), existing gasifier designs can be adapted to use any type of coal as gasifier feed. Thus, coal characteristics (and other feedstock characteristics) do not offer insurmountable obstacles to its use for the production of synthesis gas as a first step to chemicals or fuel production (Speight, 2014b).

2.3.2 Liquefaction and Carbonization

Liquefaction

Coal can also be liquefied directly, without going through the production of synthesis gas. This process is termed coal-to-liquid (CTL) and is a reasonably mature technology. The process typically uses the technique of heating under pressure (up to 470°C, 200 bar) and hydrogenation where hydrogen is added to a coal-water slurry. The slurry increases the H/C ratio to a crude oil level and removes impurities such as sulfur, nitrogen, and oxygen. Coal liquefaction has been reviewed by United Kingdom Department of Trade and Industry (1999), in a paper that gives advantages and disadvantages of the technology.

Recent developments in coal liquefaction are mostly based on the pre-World War II technologies, with the exception being a process developed by Conoco based on coal dissolution in molten zinc chloride. As a response to the crude oil crisis in the l970s, two-stage liquefaction was developed, which separated coal dissolution and hydrogenation. Experience from coal liquefaction has indicated that the areas that pose the most risk to implementation on a large scale include the handling of liquids with a high load of solids that can cause mechanical difficulties in the plant. Waste handling is also an issue because of emissions of trace contaminants, such as mercury, that need to be quantified and regulated, handling of reduced sulfur by-products, and polyaromatic hydrocarbon residues that are carcinogenic and so cannot be disposed of easily.

An advantage of coal liquefaction over gasification is the thermal efficiency of fuel production, with the former being 60–70% and the latter no higher than 55% (Research Reports International 2006). Product compositions are also different, with the direct process being skewed towards lower cetane fuels, and a higher aromatic content. The use of coal liquefaction for production of chemical feedstocks has not been discussed much in the literature, although the production of lower specific gravity hydrocarbons generated from coal gasification should be more easily be adapted to prevalent petrochemical-reaction pathways to commodity chemical production.

Coal liquefaction and the upgrading of the coal liquids are being considered as a future alternative of petroleum to produce synthetic liquid fuels due to the declining crude oil reserves and the high dependence on the foreign oil supplies. Since the late 1970s, coal liquefaction processes have been developed into integrated two-stage processes, in which coal is liquefied in the presence of hydrogen in the first stage and the products are upgraded in the second stage. Upgrading of the coal liquids is an important aspect of this approach and may determine whether such liquefaction can be economically feasible.

However, coal liquids have remained largely unacceptable as refinery feedstocks because of their high concentrations of aromatic compounds and high heteroatom and metals content. Successful upgrading process will have to achieve significant reductions in the content of the aromatic components. However, the hydrogenation of coal liquids with multiring, aromatic hydrocarbons with hydrogen is a difficult process from the technological point of view due to the stable structures of the aromatic compounds and the poor dynamic yields at low pressure and low temperature. However, liquid products from coal are generally different from those produced by petroleum refining, particularly as they can contain substantial amounts of phenols. Therefore, there will always be some question about the place of coal liquids in refining operations as well as the use of the liquids as suitable feedstocks for the production of organic chemicals.

Coprocessing of coal and a noncoal liquid hydrocarbon has also been considered, in particular for heavy oil. Coprocessed feed has the advantage that recycled solvents are not required, and is a method of facilitating refining of heavy oil.

Carbonization

Carbonization (which includes pyrolysis) is the oldest direct methods of production of liquids from coal, involving heating of the coal and capture of volatilized liquids leaving a char reduced in hydrogen. The amount of liquid generated is small, less than 20%, and the quality is poor being a complex blend of chemicals with water contamination. However, carbonization is a process predominantly for the production of a carbonaceous residue (coke) by the thermal decomposition (with simultaneous removal of distillate) of organic substances (Wilson and Wells, 1950; McNeil, 1966; Gibson and Gregory, 1971). The process, which is also referred to as destructive distillation, has been applied to a whole range of organic (carbon-containing) materials particularly natural products such as wood, sugar, and vegetable matter to produce charcoal. In this present context, the carbonaceous residue from the thermal decomposition of coal is usually referred to as “coke” (which is physically dissimilar from charcoal) and has the more familiar honeycomb-type structure. But, coal carbonization is not a process that has been designed for the production of liquids as the major products.

Carbonization is essentially a process for the production of a carbonaceous residue by thermal decomposition (with simultaneous removal of distillate) of organic substances:

CorganicCcoke/char/carbon+liquids+gases

si3_e

The process may also be referred to as destructive distillation and has been applied to a whole range of organic materials, but more particularly to natural products such as wood, sugar, and vegetable matter to produce charcoal. Coal usually yields “coke” which is physically dissimilar from charcoal and has the more familiar honeycomb-type structure.

Coal tar is the volatile material that is released during the thermal decomposition of coal and which condenses at room temperature. The tar may be composed of solid material (pitch) and liquid or semisolid materials (coal tar). The carbonization of coal to produce coal gas for street and house lighting in the closing years of the 18th century produced substantial quantities of tar which (during the following 50 years) were mostly discarded as a troublesome and unnecessary by-product. However, the development of a western European chemical industry brought increasing importance to coal tar as a source of the precursors that were to be used for the synthesis of dyes as well as raw materials for the production of solvents, pharmaceutical products, synthetic fibers, and plastics (Speight, 2013, 2014a, 2016).

Coal tar can also be upgraded to gasoline and other liquid fuels. In fact, in the manner of crude petroleum, high temperature tars can be fractionated by distillation into (1) light oil, (2) middle (or tar acid) oil, and (3) heavy (or anthracene) oil. This primary separation is carried out by means of batch stills (vertical or horizontal; 3000–8000 US gallon, 11–30 × 103 L, capacity) or by means of continuous “pipe” stills in which the tar is heated to a predetermined temperature before injection into a fractionating tower.

The light oil fraction (b.p. 220°C; 430°F, c.f. petroleum naphtha b.p. 205°C; 400°F) consists mostly of benzene (45–72% w/w), toluene (11–19% w/w), xylene (3–8% w/w), styrene (1–1.5% w/w), and indene (1–1.5% w/w) and is processed either into gasoline and aviation fuel components or is fractionated further to provide solvents and petrochemical feedstocks. In either case, upgrading involves removal of sulfur compounds, nitrogen compounds, and unsaturated materials. This is usually accomplished by acid-washing in batch agitators or by hydrogenation over a suitable catalyst (e.g., cobalt-molybdenum or nickel-tungsten on a support). Thus, in the acid wash, the crude material is mixed with strong sulfuric acid, neutralized (with ammonical liquor or caustic soda), and after separation of the aqueous phase, steam-distilled or stripped of higher molecular weight material by centrifuging. The hydrogenation process conditions vary with the nature of the material to be removed (such as removal of sulfur or removal of olefin products by hydrogenation), but could typically be 300–400°C (570–750°F) and 500–1500 psi hydrogen. The middle oil typically boils over the range 220–375°C (430–710°F) and after extraction of the tar acids, tar bases, and naphthalene, can be processed to obtain diesel fuel, kerosene, or creosote.

Coal tar creosote consists of aromatic hydrocarbons, anthracene, naphthalene, and phenanthrene derivatives. At least 75% of the coal tar creosote mixture is polycyclic aromatic hydrocarbons (PAHs). Unlike the coal tars and coal tar creosotes, coal tar pitch is a residue produced during the distillation of coal tar. The pitch is a shiny, dark brown to black residue which contains PAHs and their methyl and polymethyl derivatives, as well as heteronuclear compounds. Coal tar pitch is the tar distillation residue produced during coking operations. The grade of pitch thus produced is dependent on distillation conditions, including time and temperature. The fraction consists primarily of condensed ring aromatics, including 2–6 ring systems, with minor amounts of phenolic compounds and aromatic nitrogen bases. The number of constituents in coal tar pitch is estimated to be in the thousands.

In this context, it should be noted that the tar acids, which are mostly phenol derivative, including cresol derivatives (cresol is CH3C6H4OH) and xylenol derivatives [xylenol is (CH3)2C6H3OH] which can be recovered by mixing the crude middle oils with a dilute solution of caustic soda, separating the aqueous layer, and passing steam through it to remove residual hydrocarbons. The acidic products—the phenol derivative, including cresol derivatives (cresol is CH3C6H4OH) and xylenol derivatives—are then recovered by treatment of the aqueous extract with carbon dioxide or with dilute sulfuric acid and are then fractionated by vacuum distillation. Tar bases are isolated by treating the acid-free oil with dilute sulfuric acid and the bases are regenerated from the acid solution by addition of an excess of alkali (e.g., caustic soda or lime slurry). The mixture is then fractionated to produce pyridine, quinoline, and isoquinoline as well as other nitrogen-containing products.

u03-01-9780128044926
u03-02-9780128044926
u03-03-9780128044926

The temperature to which the distillation of the heavy oil fraction is taken depends on the type of residue pitch; that is desired but usually lies within the range 450–550°C (840–1020°F). In all cases, the distillate is an excellent source of hydrocarbons such as anthracene, phenanthrene, acenaphthene, fluorene, and chrysene. The residual coal tar pitches are complex mixtures that contain several thousand compounds (mostly condensed aromatic compounds) and may, by analogy, be likened to the vacuum residua that are produced in a petroleum refinery and represent the materials in petroleum which have boiling points in excess of 565°C (1050°F).

One important aspect of coal tar chemistry relates to the presence and structure of the nitrogen species in the tar. Initial work on tar nitrogen chemistry indicated that the nitrogen structure of tar is similar to the nitrogen structures in the parent coal. This was due to the belief that nitrogen in coal exists in tightly bound compounds and hence the most thermally stable structures in the coal during devolatilization. Furthermore, during the thermal reactions that led to tar formation these nitrogen compounds were released without rupture as part of the tar.

2.4 Chemicals From Tar Sand Bitumen

Tar sand bitumen, derived from unconventional sources such as tar sands in Canada and Venezuela has a higher aromatic content than conventional crude. The key aspect that needs to be addressed in the use of heavy oil for chemicals is the development of a ring opening catalyst to break down the polyaromatic tar compounds into smaller molecules. Oil shale in Colorado has the potential to be a large scale domestic supply of petroleum. Extraction technologies are currently being developed and tested, with the product of leading processes predicted to be sweet crude.

Unconventional sources of petroleum include the Canadian Athabasca oil sand, Venezuelan heavy oil, and oil shale in the Western United States. Historically, the relatively high cost of extraction of these hydrocarbons has been a major detriment to the use of heavy oil. These feedstocks are now more competitive due to the lowering costs of the production of synthetic crude oil and the volatility of the price of conventional crude oil.

The extraction of bitumen-derived crude oil follows several steps. The first step is the extraction of the heavy oil from the rock. Tar sands (7% bitumen by weight from the tar sands) and oil shale contain relatively low concentrations of hydrocarbons. Some loss (< 10% of bitumen) occurs in the primary extraction process. Increased mechanical breakup or froth treatment can increase the yield. Current research is focused on making the extraction of unconventional crude more sustainable with less of an environmental impact, including better handling of mine tailings and the water entrained in the bitumen as it is extracted from the tar sands.

To date there has been little published on the use of heavy oil, oil shale, or tar sands for chemicals production because these sources of oil have only recently (within the past two decades) become cost competitive with natural gas and with crude oil. In addition, the sources of unconventional crude oil are far from current petrochemical plants. Synthetic crude oil, derived from tar sands in Canada and from oil shale in Colorado, has a higher aromatic content than conventional crude oil. The key research aspect that needs to be addressed in the use of synthetic crude oil for the production of organic chemicals is the development of a ring opening catalyst to break down the polynuclear aromatic tar compounds into lower molecular weight less complex products.

2.5 Chemicals From Biomass

Thousands of years of innovations in agriculture have optimized crops for food and fiber production, but not for energy production. In general, the fermentation of sugar from crops such as corn and sugarcane will provide oxygenated organics, but these are often small volume niche chemicals with limited potential for large scale manufacture (Chang and Holtzapple, 2000; Chang et al., 2001a, b; Hammerschlag, 2006). However, some biobased chemicals that have potential for large scale manufacture include the carboxylic acid derivatives and glycol derivatives (Table 3.1).

Table 3.1

Conversion of Biomass to Chemicals

ProcessPrimary ProductsSecondary Products
GasificationHydrogen
Alcohol derivatives
NaphthaGasoline
KeroseneDiesel fuel
Olefin derivativesPolymers
Oxo chemicals
Synthetic natural gasHydrogen
Synthesis gas
PyrolysisHydrogen
BiooilVarious chemicals
Anaerobic digestionBiogasHydrogen
FermentationEthanolEthanol derivatives
Hydrogen

Projects that can lead to the production of organic chemicals from various sources include: (1) biomass gasification, (2) fermentation of sugars, (3) decomposition of cellulose, (4) separation of lignin and other plant components, (5) high temperature pyrolysis, and (6) biorefining of wood and waste materials (Speight, 2011c). Issues in the replacement of petroleum by biomass feedstocks include impurities, variabilities of feedstock composition, distributed supply, scalability, and pathways for breakdown of cellulose. Although some large-scale chemicals production occurs as a by-product of fuel production, widespread use of biomass feedstocks for commodity chemical manufacture will require sustained research and development in a variety of fields such as plant science, microbiology, genomics, catalysis, and chemical separation technologies.

Chemicals can be manufactured from biomass through gasification, pyrolysis, and fermentation in dedicated plants or in biorefineries. In general, the fermentation of sugar from crops such as corn and sugarcane will provide oxygenated organics, but these are often small volume niche chemicals with limited potential for large scale manufacture. By-products of transportation fuel production, such as biodiesel, may be used in bulk chemical manufacture—one example being glycerin (also called glycerol, glycerine, and propane triol):

u03-04-9780128044926

Efficient utilization of intermediate reagents such as these will eventually lead to the development of new synthesis pathways, or optimization of synthetic routes that are currently only being done in the laboratory.

Use of biomass to make biofuels such as ethanol and biodiesel, is of great interest in the United States and elsewhere (Speight, 2011c). Biofuels provide energy security, support local agriculture, and reduce net emissions of carbon dioxide to the atmosphere. The use of biomass to make chemicals has received less public attention to date. Chemical manufacturers currently use biomass to make organic acids, textile fibers, polymers, adhesives, lubricants and greases, and soy-based inks.

Biorefining is the name given to the use of biologically derived feedstocks for chemicals manufacture. The biorefinery concept generally involves feeding biomaterials, along with waste oils and other carbon-based materials, into steam or catalyst crackers to make chemicals (Speight, 2011c). Alternatively, these feedstocks may be hydroprocessed directly. Mills are used to process biomass (corn, pulp) to produce carbohydrates, oils, lignin, and fuel compounds. Once broken down, fermentation will produce alcohols from sugars and starch. Biorefining feedstocks are crops, waste plant or animal material, and recycled fibers. Eventually, biorefining will not only utilize the starch or sugar component of biomaterials, but also consume lignin, hemicellulose, and cellulose in value-added processes beyond the current practice of burning these materials for fuel. An example is bioconversion of sugar derivatives to produce polyols (polyhydroxy compounds which are building block chemicals). In addition, another example of biorefining is the conversion of vegetable oils to lubricants, hydraulic fluids, and other chemical products that can serve as monomer for polymerization processes or building block chemicals.

Bioprocessing of corn to produce ethanol can either be done by wet-milling or dry-milling—the wet-milling process produces gluten feed, gluten meal, and corn oil, as well as ethanol while the dry-milling process gives ethanol only. In spite of additional expenses incurred for separations equipment and operation, these by-products can be sold to improve the economics of ethanol production. Wet mills are large, and generally require a coal-fired plant to operate, whereas dry mills (~ 40 M gal) are smaller, and use natural gas. The by-product of sugarcane processing is bagasse, which can be burned for fuel, providing sufficient energy to run the processing plant. Also, the production of biodiesel (fatty acid methyl esters) from various feedstocks as well as from biomass residues is of interest in many countries, particularly in the developing world. The processes for manufacturing diesel is an alternative to fermentation and the use of the reactive distillation concept to produce esters of dibasic acids takes advantage of differences in volatility to give high purity products. The reactive distillation is a process where the still pot is the chemical reactor and separation of the product from the reaction mixture does not need a separate distillation step, thereby saving process energy (for heating) as well as the need from an extra piece of equipment (the reactor). Thus, production of biofuels will have a direct impact on the chemicals industry. For example, glycerine produced as by-product of biodiesel manufacture, will supplant propylene as a feedstock in the production of epichlorohydrin which can serve as an intermediate organic chemical for the production of epoxy resins.

u03-05-9780128044926

Epichlorohydrin (ECH) is an organochlorine compound and an epoxide (a three-membered ring that contains two carbon atoms and an oxygen atom) and, despite the name, it is not an acid chloride (like acetyl chloride, CH3COCl):

u03-06-9780128044926
An acid chloride; when Rglyph_dbndCH3, the acid chloride is acetyl chloride.

2.5.1 Thermochemical Gasification

Thermochemical gasification is one of several thermochemical conversion processes which also include combustion and pyrolysis and serve to convert carbonaceous feedstock to gases, liquids, and solid products.

Combustion is the complete conversion (in the presence of oxygen) to carbon dioxide, water, and energy. Combustion is also a means of converting waste to energy and involves oxidation of the fuel for the production of heat at elevated temperatures without generating useful intermediate fuel gases, liquids, or solids. Combustion typically employs an excess of the oxidizer (air) to ensure maximum fuel conversion and the products of combustion processes include heat, oxidized species (such as carbon dioxide and water), products of incomplete combustion (such as carbon monoxide and hydrocarbons), other reaction products (most as pollutants), and ash from any inorganic mineral matter (or organo-minerals) in the feedstock.

Pyrolysis is the thermal degradation of a material usually without the addition of any air or oxygen. The process is similar to gasification but generally optimized for the production of fuel liquids or pyrolysis oils (sometimes called biooils if biomass feedstock is used). Pyrolysis also produces gases and a solid char product. Pyrolysis liquids can be used directly (e.g., as boiler fuel and in some stationary engines) or refined for higher quality uses such as organic chemicals, gasoline, diesel fuel, and other products.

Gasification typically refers to conversion in an oxygen- or air-deficient environment to produce fuel gases (such as synthesis gas). The fuel gases are principally carbon monoxide, hydrogen, methane, and lighter hydrocarbons, but depending on the process used, can contain significant amounts of carbon dioxide and nitrogen, the latter mostly from air. Gasification processes also produce liquid products (tars, oils, and other condensates) and solids (char, ash) from solid feedstocks. The combustion of gasification-derived fuel gases generates the same categories of products as direct combustion of solids, but pollution control and conversion efficiencies may be improved. Synthesis gases can produce fuel products and other chemicals by chemical reactions such as Fischer-Tropsch synthesis (Chadeesingh, 2011).

Synthesis gas for commodity chemical production can be derived from any carbonaceous feedstock, such as coal, petroleum residua, or biomass. Issues such as the production of clean syngas via biomass thermochemical processing are similar to issues associated with coal gasification. Gasification of biomass can take place under slightly milder conditions than coal gasification (800–1000°C at 20–30 bar instead of 1400°C at 20–70 bar). Biomass (feedlot and chicken litter) can also be combined with a coal-syngas feedstock (Priyadarsan et al., 2004, 2005).

However, using biomass for thermochemical conversion raised several issues and some pretreatment is necessary and is unique to gasification of biomass. Biomass has a large water content that must be removed before gasification. Also, biomass components (alkali metals, halides, sulfur compounds, and tars) have a significant potential to poison downstream noble metal catalysts used in production of syngas and chemicals (Ragauskas et al., 2006). Technologies have been developed to handle these impurities, but they add to the complexity and cost of the gasification process. In addition, because production of biomass requires a large land base, feedstocks are diffuse, and manufacturing is distributed (e.g., forest pulp mills). Hence, additional methods as required by the character of the feedstock, such as feedstock densification or on-site drying prior to shipping, will be required to achieve economies of scale.

2.5.2 Sugar Fermentation

Fermentation is a metabolic process that converts sugar to acid derivatives, various gases, or alcohol. Another definition is that fermentation is the chemical process by which molecules such as glucose (C6H12O6) are decomposed anaerobically. The process can involve complete decomposition of the glucose to carbon dioxide and water (+ energy) or can be adapted to produce ethanol (ethyl alcohol + energy):

C6H12O6fermentationCO2+H2O+energyC6H12O6fermentationC2H5OH+energy

si4_e

The presence of oxygen during the fermentation offers routes to a suite of other chemical products.

Thus, fermentation offers a pathway for the use of biomass as a means of producing organic chemicals either directly or organic chemicals as feedstocks for the production of other chemicals. In fact, the fermentation of sugars has been the focus of much recent investigation because of the chemical character and properties of the products, which include: (1) the presence of functional groups that give rise to building block chemicals, (2) the facile adaptation to petrochemical pathways, and (3) the performance of the production process, which is a well-tested operation.

Chemical transformation of sugars can take place by oxidation (such as during the decomposition of polysaccharides of which starch is an example), oxidative dehydration (of the six-carbon sugars), hydrogenation (sometimes acid-catalyzed) of cellulose derivatives and sugar derivatives, acid amination, and esterification of oils. These products can be further converted by chemical means to derivatives by: (1) oxidation, which tends to be less important for biomass as these compounds are already oxidized, (2) by hydrogenation, (3) by dehydration, (4) by bond cleavage, and (5) by direct polymerization. Biological reactions to derivatives can also be achieved with the advantages that they are generally enzymatic and so very selective, and may go directly from sugar to the end product.

2.5.3 Nonsugar Fermentation

Annual crops such as corn that have a high sugar yield are typically difficult to grow, and need fertilization, irrigation, herbicides, and pesticides. In addition, these crops as well as waste biomass have been engineered for food production, rather than for efficiency of photosynthesis, or energy, or chemicals production (Ragauskas et al., 2006). Use of cellulose from waste biomass would allow raw material to come from corn stalks (Thanakoses et al., 2003a), wood, bagasse from sugarcane (Thanakoses et al., 2003b), and even animal products (if proteins can be used) as well as various types of waste (Chan and Holtzapple, 2003; Aiello-Mazzarri et al., 2005, 2006).

Conversion of cellulose to fuel and hydrocarbons is a multistage process. The biomass must first be physically or chemically broken down, to separate the cellulose from other components, such as lignin. Pretreatment issues dominate cellulose and lignin processing, and often involve acid- or base-catalyzed hydrolysis to facilitate enzymatic breakdown. The costs of pretreatment are high, but some suggest that few major technical improvements in the chemical or acid processing are possible. However, pretreatment is an active area of research, one example being investigation of improvements in lime pretreatment (e.g., Chang and Holtzapple, 2000; Chang et al., 2001a, b).

Once isolated, the next step is the breakdown of cellulose to form sugars. The natural rotting process, facilitated by bacteria or fungi, is slow. Enzymatic hydrolysis increases the rate of the process, but the production of by-products is a problem when the processes of oxidation produce aldehydes and acids as well as sugars. Once produced, these sugars can be fermented and processed by conventional means. Hence, the key needs for fermentation of biomass include bacteria that break down cellulose quickly. Other process goals include the use of different conversion pathways (Lynd, 1996). In addition, separation processes are essential in handling of the varied and variable biofeedstocks which includes separation of cellulose from lignin and other plant materials and separation of the by-products from the product after fermentation is complete. A separation step in itself may allow production of a value-added chemical, such as the production of xylitol [CH2OH(CHOH)3CH2OH, a polyalcohol or sugar alcohol] from the pretreatment of cellulose.

2.5.4 Pyrolysis

A third means of extracting chemicals from biomass feedstocks is by means of pyrolysis (Speight, 2011c). The biomass feedstocks can be wood wastes, bark, or other forest products and the resulting products are biooils which are comprised of oxygenated organic compounds, and water. Pyrolysis is complex, incorporating both evaporation and combustion of a chemically unknown fuel. The process involves (1) heating and vaporization of water, (2) separation of the volatile components, and (3) formation of a porous char or cenosphere from the high-boiling nonvolatile components. Unlike direct combustion, pyrolysis occurs at high temperatures, on the order of 400°C (750°F).

High temperature pyrolysis has been suggested as a more effective method of black liquor gasification than traditional lower temperature approaches because tarry residues are less of a problem. Black liquor is the waste product from the Kraft process when digesting pulpwood into paper pulp, removing lignin, hemicellulose derivatives, and other extractives from the wood to free the cellulose fibers. At temperatures in the range of 700°C and 1000°C (1290–1830°F) the process yields tar products, semivolatile products, and nonvolatile char that were identified from the heating of black liquor (Sricharoenchaikul et al., 2002). In addition, in organic chemicals such as single benzene ring derivatives up to pyrene and fluoranthene (4-ring compounds) derivatives are produced.

u03-07-9780128044926
u03-08-9780128044926

3 Process Chemistry

Industrial organic chemicals are those organic compounds produced in the United States and other countries as single compounds or as the more complex mixtures, such as fuels that are composed of organic compounds (Tables 3.2 and 3.3) (Kroschwitz et al., 1991; Chenier, 1992; Speight, 2002, 2014a, b; Wittkoff et al., 2012). Historically, during the 19th century and early 20th century most organic chemicals had been obtained as by-products from the coking of coal, e.g., from coal oil. During the 20th century, however, oil and natural gas became the dominant sources of the world's industrial organic chemicals. By 1950, a substantial portion of the industrial organic chemicals were being produced from crude oil and natural gas, and by 2000 more than 90% of the organic chemical industry was based on crude oil and natural gas as the prime feedstocks. As a result, the term petrochemicals has almost become synonymous with industrial organic chemistry and the chemicals industry.

Table 3.2

Example of Organic Chemicals Manufactured in the United States

ChemicalVolume of Production 106 tons (US, 1997)Alternative Feedstock or PrecursorIndustry/Product
Olefins
Ethylene24.125Gasification of coal, biomass. DME
Methanol (MTO. MTP)
Gas to polymers
Propylene14.350
Butadiene (1,3-)2.038
Ammonia14.204Gasification of coal, biomass
Natural gas, Haber cycle
Fertilizer, reagent, explosives
Chlorinated organics
Ethylene dichloride10.088Chlorination of olefinsPolyurethanes, solvents, pulp and paper solvents
Vinyl chloride8.753
Methyl chloride0.563
MTBE9.038CH3OH + isobutyleneFuel additive
Aromatics
Benzene7.463Gasification of coal, biomass
From methanol from GTL
Friedel-Crafts alleviation
Polymerization/surfactants
Ethylbenzene6.950
Toluene4.138
p-Xylene3.963
Cumene2.913
Methanol6.013Gasification of coal, biomass
FT
Building block
Urea5.918Ammonia, CO2;Fertilizer, resins, adhesives
Styrene5.700EthylbenzenePolystyrene
Terephthalic acid5.000p-XyleneIntermediate
Aldehydes
Formaldehyde4.188From methanol (natural gas)Building block for olefins
Oxygenated organics
Ethylene oxide3.550Gasification of coal, biomass, through ethylene oxideBuilding block, reagent
Propylene oxide1.963
Ethylene glycol2.813Hydration of ethylene oxidePolyester
Propylene glycol0.538
Carboxylic acids
Acetic acid and anhydride2.425Gasification of coal, biomass
Methanol from syngas
Adds C2, reagents and intermediates
Phenol2.175CumeneResins, paints, adhesives, coatings, solvents, polycarbonate
Bis phenol A0.863
Acrylonitrile1.663NH3 + propylene, propaneAcrylic fibers
Esters
Vinyl acetate1.500From ethyleneIntermediate coatings, plastics
Methyl methacrylate0.313
Acetone1.463Gasification of coal, biomass, BTEXIntermediate
Cyclohexane1.100Hydrogenation of benzeneIntermediate
Caprolactam0.825Oxidation of cyclohexaneNylon
Aniline0.713Nitro or chlorobenzeneIntermediate
Isopropyl alcohol
Butanol
0.700Gasification of coal, biomass, propyleneIntermediate

t0015

Table 3.3

Manufacture of Fuels

Paraffins: Ethane Propane ButanesGasification of coal, biomass Methanol dehydrogenation FT liquidFuel building block
EthanolFermentation of biomass Gasification of coal, biomassFuel
Biodiesel, GlycerolEnzymatic transformation of biomass Transesterification of oilsFuel Building block Polyurethane, glycol, 1-propane diol, biodiesel
Dimethyl etherReplacement for propane

Nevertheless, terminology aside, there are several types of process reactions common to the organic chemicals manufacturing industry: (1) polymerization, (2) oxidation, and (3) addition. Polymerization is a chemical reaction usually carried out with a catalyst, heat, or light (often under high pressure) in which a large number of relatively simple olefin derivatives (such as ethylene, CH2glyph_dbndCH2) combine to form a high-molecular weight product [a macromolecule, such as polyethylene, Hglyph_sbnd(CH2CH2)nglyph_sbndH]. Oxidation, in the strictest chemical sense, is the combination of oxygen chemically with another substance although this name is also applied to reactions where electrons are transferred (Chapter 2). The addition reaction covers a wide range of reactions where a double bond (> Cglyph_dbndC <) or a triple bond (glyph_sbndCglyph_tbndCglyph_sbnd) is broken and a component molecule is added to the reactant molecule. The alkylation reaction (often considered to be an addition reaction as practiced in the petroleum refining industry) in which a hydrogen atom is converted to an alky group can be considered an addition reaction:

u03-09-9780128044926
Alkylation by replacement of an active hydrogen by an alkyl group.

However, an important aspect of process chemistry is that most industrial organic chemistry is considered to fall into one of the following categories: (1) C-1 chemistry, based on synthesis gas, also called syngas which is a variable mixture of carbon monoxide (CO) and hydrogen (H2) produced by the high temperature reaction of steam water with crude oil residua, coal, oil, or natural gas, or for that matter any carbonaceous material, including biomass (Speight, 2013, 2014a,b), (2) C-2 chemistry, based on the chemistry of ethylene, (3) C-3 chemistry, based on the chemistry of propylene, (4) C-4 chemistry, based on the chemistry of the butane isomers and the butene isomers, and (5) BTX chemistry, based on the chemistry of benzene, toluene, and the xylene isomers.

Alkenes or olefins (ethylene, propylene, butene isomers, and butadiene) are mainly produced via thermal steam cracking as used in the petroleum refining industry (Speight, 2014a, 2016). In the steam cracking process, a fraction of crude oil (usually isolated by distillation without any further purification, unless a catalyst is involved) is mixed with water and heated using a short residence time (usually on the order of 1 s) at 800–900°C (1472–1652°F). These process parameters cause rupture of the carbon-carbon bonds to yield lower molecular weight products which also are reduced in hydrogen content compared to the feedstock (as deduced by comparing the atomic hydrogen-to-carbon ratio of the feedstock and products) with the formation of double bonds. On the other hand, the BTX compounds (benzene, toluene, and xylene isomers) the simplest aromatics, are largely produced during catalytic reforming (the Platforming process). In this process a naphtha fraction that is rich in the pentane t-nonane alkane hydrocarbons (C5 to C9 alkanes) is reacted at temperatures on the order of 450°C (842°F) and approximately 300–400 psi, over a platinum (Pt/SiO2) catalyst, to yield a reaction product that contains approximately 60% aromatic hydrocarbons. Typically, the product mixture might contain approximately 3% v/v benzene, 12% v/v toluene, 18% v/v xylene isomers, and 27% v/v C9 alkylbenzenes. The product, because of the presence of the aromatic components has a high octane number and can be blended (as a mixture) as individual aromatic constituents as a blend-stock for gasoline production. Because benzene is much more in demand for industrial purposes than toluene, the methyl group of toluene is often removed by hydrogenation.

C6H5CH3Toluene+H2C6H6Benzene+CH4

si5_e

Thus, the following sections relate to the production of high-volume organic chemicals which also serve to illustrate three key points, which are: (1) primary building blocks are typically used in more reactions than the chemical building blocks in the later stages of chemical production, (2) most feedstocks for chemicals production can participate in more than one reaction, and (3) there is typically more than one reaction route to an end-product. Furthermore, the end-products of all of these chemical processes can be used in several commercial applications.

3.1 C-1 Chemistry

C-1 chemistry refers to the conversion of carbon-containing organic chemicals materials that contain one carbon atom per molecule into more valuable products. The feedstocks for C-1 chemistry include natural gas (predominantly methane), carbon dioxide, carbon monoxide, methanol, and synthesis gas (syngas, a mixture of carbon monoxide and hydrogen). Synthesis gas is produced primarily by the reaction of natural gas, which is principally methane, with steam as well as by the reaction of any high-molecular weight carbonaceous feedstock (such as coal, petroleum residua, petroleum coke, or biomass) (Speight, 2013, 2014a,b). The availability of synthesis gas from such feedstocks coal gasification is expected to increase significantly in the future because of increasing development of IGCC power generation and it is anticipated that the refinery of the future will also have a gasifier integrated into the refinery unit processes (Speight, 2011b).

Synthesis gas is also the feedstock for all methanol and Fischer-Tropsch plants. In fact, many important organic chemicals can be produced from synthesis gas (syngas)—the carbon monoxide (CO) and hydrogen (H2) mixture produced as a result of the gasification of a variety of carbonaceous feedstocks either as a single feedstock or as a mixture of feedstocks (Speight, 2014b). They range from simple molecules, such as methanol, to high-grade synthetic crude oil. The basic reaction for conversion of synthesis gas to mixtures of hydrocarbons (Fischer-Tropsch reaction) was used in Germany during World War II to produce fuel mixtures for the military diesel and gasoline engines. Since the 1950s South Africa has also used this reaction, and currently there is much interest in using it to convert natural gas (methane) to more easily transported liquids as well as for the end use of these liquids.

Ammonia (NH3), although it is not an organic chemical, is often considered as part of C-1 chemistry, since it is produced via a reaction that uses hydrogen gas obtained from methane. It is made by the Haber process. Ammonia and its derivatives, HNO3, NH4NO3, and CO(NH2)2, are key fertilizers and ingredients for explosives, and their production consumes nearly 5% v/v of the worldwide natural gas.

Methanol (methyl alcohol, CH3OH), an important C-1 chemical that is used as a solvent as well as a precursor for many organic chemicals, is made by the hydrogenation of carbon monoxide—a process developed in the 1920s:

CO+2H2CH3OH

si6_e

For example, as a precursor to other chemicals, a major use of methanol is the production of acetic acid. Acetic acid (ethanoic acid, CH3CO2H) was for many years made by the oxidation of ethanol:

C2H5OH+O2CH3COOH+H2O

si7_e

However, there are other processes by which acetic acid can be manufactured including the carbonylation of methanol:

CH3OH+COCH3CO2H

si8_e

C-1 chemistry is expected to become a major means of chemical production (or, at least, an initial step in the production of many organic chemicals) for the production of chemicals and transportation fuels in the near future (Speight, 2011b). In addition, synthesis gas is a major source of refinery hydrogen for use in a variety of refining processes (i.e., processes that require a hydrogen presence in the reactor) as well as for a variety of chemical and petrochemical operations.

3.2 C-2 Chemistry

C-2 chemistry usually refers to the processes that use ethylene as the starting organic chemical and it is one of the largest volume organic compounds used worldwide. Ethylene is produced in the petrochemical industry by steam cracking. In this process, gaseous or light liquid hydrocarbons are heated to 750–950°C (1380–1740°F). The products are quenched to prevent further reactions (often leading to the formation of tar-like products) and ethylene is separated from the resulting mixture by repeated compression and distillation. In a related process used in oil refineries, high-molecular weight feedstocks are cracked over zeolite catalysts. Feedstocks such as naphtha and gas oil require at least two quench towers downstream of the cracking furnaces to recirculate pyrolysis-derived gasoline and process water. When cracking a mixture of ethane and propane, only one water quench tower is required.

The major uses for ethylene are in the synthesis of polymers (such as polyethylene) and for the production of ethylene dichloride, a precursor to vinyl chloride (Tables 3.4 and 3.5). Other important products are ethylene oxide (a precursor to ethylene glycol) and ethylbenzene (a precursor to styrene).

Table 3.4

Industrial Uses of Ethylene

ProcessTarget ProductProcess ConditionsReaction ComponentsOther Characteristics
Pressure (MPa)Temperature (°C)Catalyst
PolymerizationLow-density polyethylene (LDPE)60–350350Oxygen or peroxide
High-density polyethylene0.1–2050–300Molybdenum Chromium oxide
PolyethyleneLowAluminum alkyls Titanium oxide
OxidationEthylene oxide1–2250–300Silver1,2-Dichloro-e thane, oxygen60% is converted to ethylene glycol using an acid catalyst
Acetaldehyde0.3120–130Copper chloride/palladium chlorideOxygenVapor phase
Vinyl acetate0.4–1170–200PalladiumAcetic acid
Addition
HalogenationhydrohalogenationEthylene dichloride60Iron, aluminum, copper, or antimony chloridesChlorineFeedstock for vinyl chloride and trichloroethylene and tetrachloroethylene
Ethyl chloride0.3–0.5Aluminum or iron chloridesHClPrecursor of styrene
AlkylationEthyl benzeneAluminum, iron, and boron chloridesBenzene
HydroformationPropionaldehyde4–3560–200CobaltSynthesis gas (carbon monoxide and hydrogen)

t0025

Table 3.5

Industrial Uses of Vinyl Chloride

ProcessTarget ProductProcess ConditionsReaction ComponentsOther Characteristics
Pressure (MPa)Temperature (°C)Catalyst
PolymerizationPolyvinylchloride50Peroxides
Substitution at the carbon-chlorine bondVinyl acetates, alcholates, vinyl esters and vinyl ethersPalladiumAlkyl halides
AdditionVarious halogen addition products

t0030

Ethylene oxide (C2H4O, also called oxirane) is a cyclic ether that is a colorless flammable gas at room temperature, with a faintly sweet odor. It is the simplest epoxide—a three-membered ring consisting of one oxygen atom and two carbon atoms:

u03-10-9780128044926

Because of this molecular structure, ethylene oxide easily participates in addition reactions, such as ring opening followed by further reactions and polymerization. This chemical reactivity has made ethylene oxide a key industrial chemical and is usually produced by direct oxidation of ethylene in the presence of a catalyst. It is extremely flammable and explosive and is used as a main component of thermobaric weapons—therefore, it is commonly handled and shipped as a refrigerated liquid. While ethylene itself is not generally considered a health threat, ethylene oxide has been shown to cause cancer.

Most ethylene oxide (about 60% v/v) is converted to ethylene glycol (HOCH2CH2OH) via acid-catalyzed hydrolysis:

CH2CH2+2OHHOCH2CH2OH

si9_e

Caution is always advised when handling a toxic chemical such as ethylene glycol. It is a toxic chemical that is used as an antifreeze, heat transfer agent, and also as a solvent in industrial organic chemical facilities. Long-term inhalation exposure to low levels of ethylene glycol may cause throat irritation, headache, and backache while exposure to high concentrations can lead to loss of consciousness. Liquid ethylene glycol is irritating to the eyes and skin and toxic effects from ingestion of ethylene glycol include damage to the central nervous system and kidneys, intoxication, conjunctivitis, nausea and vomiting, abdominal pain, weakness, low blood oxygen, tremors, convulsions, respiratory failure, and coma. Ethylene glycol readily biodegrades in water and in soils—biodegradation is probably the dominant removal mechanism.

Vinyl chloride (CH2glyph_dbndCHCl) is the second-largest-volume chemical made from ethylene and is produced by adding chlorine to ethylene (to produce ethylene dichloride, ClCH2CH2Cl) and then thermal treatment to remove hydrogen chloride from the intermediate ethylene dichloride.

CH2CH2+Cl2ClCH2CH2ClClCH2CH2ClCH2CHCl+HCl

si10_e

The uses of vinyl chloride include polymerization to polyvinyl chloride (PVC) which is used to make pipe, floor covering, wire coating, house siding, imitation leather, and many other products (Table 3.5). Vinyl chloride is one of the largest commodity chemicals but is also considered a human carcinogen by the US EPA. Vinyl chloride polymers are the primary end use but various vinyl ethers, esters, and halogen products can also be made as shown in the table below.

Styrene (phenylethylene, vinyl benzene, C6H5CHglyph_dbndCH2) is made from ethylene by reaction with benzene to form ethylbenzene, followed by dehydrogenation.

C6H6+CH2CH2C6H5CH2CH3C6H5CH2CH3C6H5CHCH2+H2

si11_e

Over 50% of manufactured styrene is polymerized to polystyrene for toys, cups, containers, and foamed materials used for insulation and packing. The rest is used to make styrene copolymers, such as styrene-butadiene rubber (SBR).

At the high-molecular weight end of the uses of ethylene, polyethylene comes in two basic types: high density and low density. The original polymer was a highly flexible branched product, first prepared in 1932 by a process that required high temperatures and ultrahigh pressures. It is now known as low-density polyethylene (LDPE), to differentiate it from a linear polymer discovered later and known as high-density polyethylene (HDPE). For many applications the original branched LDPE has now been replaced by linear low-density polyethylene (LLDPE). HDPE is more rigid and less translucent than LDPE or LLDPE, and it has a higher softening point and tensile strength. HDPE is used to make bottles, toys, kitchenware, and so on, whereas LDPE and LLDPE are predominantly used for film used in packaging (such as plastic bags).

3.3 C-3 Chemistry

C-3 chemistry is based on the chemistry of propylene (CH3CHglyph_dbndCH2) from which polypropylene is produced. At room temperature and atmospheric pressure, propylene is a gas, and as with many other alkenes, it is also colorless. Propylene is produced in nature as a by-product of vegetation and fermentation processes. On the industrial scale, propylene is produced from crude oil and natural gas—during crude oil refining ethylene, propylene, and other compounds are produced as a result of the thermal decomposition (cracking) of higher molecular weight larger hydrocarbons (Speight, 2014a, 2016). Thus, a major source of propylene is from naphtha cracking which is a major process for ethylene production, but propylene is also a product from refinery cracking processes that produce other products. Propene can be separated from product mixtures by fractional distillation.

Propylene is also produced by olefin disproportionation—a reversible reaction between ethylene and butenes in which double bonds are broken and then reformed to form propene, for example in the simplest chemical sense the reaction can be represented as:

CH2CH2+CH3CH2CHCH22CH3CHCH2

si12_e

Also, propylene is produced by dehydrogenation of propane converts propane into propene and by-product hydrogen:

CH3CH2CH3CH3CHCH2+H2

si13_e

In addition, olefins cracking processes include a broad range of technologies that catalytically convert higher molecular weight olefins (butene olefins to octene olefins, C4 to C8) into mostly propene (predominantly) and ethylene (lesser amounts).

The primary products manufactured from propylene are polypropylene, acrylonitrile, propylene oxide, and isopropyl alcohol (Table 3.6). Acrylonitrile and propylene oxide have both been shown to cause cancer, while propylene itself is not generally considered a health threat. Polypropylene is used to make injection-molded articles, such as automotive battery cases, steering wheels, outdoor chairs, toys, and luggage as well as fibers for upholstery, carpets, and special sports clothing. Oligomers (dimers, trimers, and tetramers) of propylene, which are made by acid-catalyzed polymerization, form mixtures known as polygas, used as high-octane motor fuel.

Table 3.6

Industrial Uses of Propylene

ProcessTarget ProductProcess ConditionsReaction ComponentsOther Characteristics
Pressure (MPa)Temperature (°C)Catalyst
PolymerizationPolypropyleneAluminum alkyls/Titanium oxide
OxidationAcrylonitrile400PhosphomolybdateAmmonia
Oxygen
Commercially valuable by-products are acetonitrile and hydrogen cyanide
Propylene oxideOxygen
Ethylbenzene
Commercially valuable by-product is tert-butyl alcohol
Addition
ChlorohydrinationPropylene oxide2537TungstenHypochlorous acid
HydrolysisIsopropyl alcohol267Water

t0035

Propylene oxide is made via several methods. The classical one involves treating propylene with chlorine water to produce propylene chlorohydrin, and then using base to split out HCl. The primary use for propylene oxide is its oligomerization (to polypropylene glycols). These products combine with diisocyanates to produce high-molecular weight polyurethane foams, which make very good padding for furniture and vehicle seats. Manufacture of propylene glycol (CH3CHOHCH2OH) consumes about 30% of the propylene oxide produced. Like ethylene oxide, propylene oxide undergoes hydrolysis to yield the corresponding glycol. Propylene glycol is mainly used to make polyester resins, but it is also used in foods, pharmaceuticals, and cosmetics.

Another chemical, acrylonitrile (CH2glyph_dbndCHCglyph_tbndN) is made by direct ammoxidation of propylene:

2CH3CHCH2+2NH3+3O22CH2CHCN+6H2O

si14_e

The major use is in making polyacrylonitrile, which is mainly converted to fibers (Orlon). It is also copolymerized with butadiene and styrene to produce high impact plastics.

3.4 C-4 Chemistry

C-4 Chemistry is the chemistry of butane isomers, butylene isomers, and butadiene. Butane (C4H10) is an alkane which is a gas at room temperature and atmospheric pressure and the term includes either of two structural isomers: n-butane or isobutane (or 2-methylpropane), or to a mixture of these isomers. However, using the correct nomenclature, butane refers only to the n-butane isomer:

u03-11-9780128044926

Whatever the structure, butane isomers are highly flammable, colorless, and easily liquefiable gases.

Butene (C4H8) is a colorless gas that is present in crude oil as a minor constituent in quantities that are too small for viable extraction. Typically, butene is obtained by catalytic cracking of higher molecular weight hydrocarbons produced during crude oil refining. The cracking process yields a mixture of products, and the butene is extracted from this by fractional distillation. Butene can be used as the monomer for polybutene but polybutene is therefore commonly used as a copolymer (mixed with another polymer, either during or after reaction), such as in hot-melt adhesives.

Among the molecules which have the chemical formula C4H8 four isomers are alkenes which have four carbon atoms and one double bond in the structure but have different chemical structures. The IUPAC names and the common names of the butene isomers are:

u03-12-9780128044926

All four of these isomers are gases at room temperature and pressure but can be liquefied by lowering the temperature or raising the pressure. These gases are colorless, but do have distinct odors, and are highly flammable. Although not naturally present in crude oil in high percentages, they can be produced in a refinery by catalytic cracking or from petrochemical intermediates. Although they are stable compounds, the carbon-carbon double bonds make them more reactive than the respective alkane hydrocarbons.

Because of the double bonds, these 4-carbon alkenes can act as monomers in the formation of polymers, as well as having other uses as petrochemical feedstocks. They are used in the production of synthetic rubber. But-1-ene is a linear or normal alpha-olefin and isobutylene is a branched alpha-olefin. In a rather low percentage, but-1-ene is used as one of the comonomers, along with other alpha-olefins, in the production of HDPE and LLDPE. Butyl rubber is made by polymerization of isobutylene with isoprene 2–7% by weight. Isobutylene is also used for the production of methyl tert-butyl ether (MTBE) and isooctane, both of which improve the combustion performance of gasoline.

1,3-Butadiene is a simple conjugated diene with the formula C4H6 and is an important industrial chemical that is used as a monomer in the production of synthetic rubber. In the simplest sense, the molecule is two vinyl groups (CH2glyph_dbndCHglyph_sbnd) joined together and the term butadiene typically refers to 1,3-butadiene which has the structure H2Cglyph_dbndCHglyph_sbndCHglyph_dbndCH2:

u03-13-9780128044926

The name butadiene can also refer to the isomer, 1,2-butadiene (H2Cglyph_dbndCglyph_dbndCHglyph_sbndCH3, also called buta-1,2-diene and methyl-allene) which is a cumulated diene but this allene is difficult to prepare and has no industrial significance. Briefly, an allene is a compound where one carbon atom has double bonds with each of its two adjacent carbon neighbors—allene derivatives are classified as polyenes—and will not undergo the same reactions as 1,3-butadiene.

Maleic anhydride is the main chemical made from n-butane. A complex catalyst is used for the oxidation reaction. The major uses for maleic anhydride are the making of unsaturated polyester resins (by reaction with glycol and phthalic anhydride) and tetrahydrofuran (by hydrogenation). MTBE is one of the leading chemicals currently being made from isobutylene (methyl propene) via the acid-catalyzed addition of methyl alcohol. MTBE has been added to gasoline as a required oxygenate. However, it is under attack as a groundwater contaminant and has been phased out of general use.

Polyisobutylene derivatives are easily made via the acid-catalyzed polymerization of isobutylene. The low molecular weight polymers are used as additives for gasoline and lubricating oils, whereas higher molecular weight polymers are used as adhesives, sealants, caulks, and protective insulation. Butyl rubber is made by polymerizing isobutylene with a small quantity of isoprene. Its main uses are in the making of truck tire inner tubes, inner coatings for tubeless tires, and automobile motor mounts. Hexamethylenediamine [HMDA, H2H(CH2)6NH2] is the principal industrial chemical made from butadiene. HMDA is polymerized with adipic acid to make a kind of nylon.

SBR accounts for about 40% of the total consumption of butadiene. SBR is the material used to make most automobile tires. Other synthetic rubbers, such as polybutadiene and polychloroprene (neoprene), make up another 25% of the butadiene market. The acrylonitrile-butadiene-styrene resin (ABS resin) is a widely used terpolymer that accounts for about 8% of the butadiene market.

3.5 BTX Chemistry

The heavy chemical industry, in its classical form, was based on inorganic chemistry, concerned with all the elements except carbon and their compounds, but including, as has been seen, the carbonates. Similarly, the light chemical industry uses organic chemistry, concerned with certain compounds of carbon such as the hydrocarbons, combinations of hydrogen and carbon. In the late 1960s the phrase “heavy organic chemicals” came into use for compounds such as benzene, phenol, ethylene, and vinyl chloride. Benzene and phenol are related chemically, and they are also related to toluene and the xylenes, which can be considered together as part of the aromatic group of organic chemicals, the aromatic compounds being most easily defined as those with chemical properties like that of benzene, toluene, and the xylene isomers.

In the crude oil refining and petrochemical industries, the acronym BTX refers to mixtures of benzene, toluene, and the three xylene (1,2-dimethylbenzene, 1,3-dimethylbenzene, and 1,4-dimethylbenzene), all of which are aromatic hydrocarbons.

u03-14-9780128044926
Benzene, toluene, and the isomeric xylenes.

If ethylbenzene (C6H5C2H5) is included in the mixture, its presence in the mixture is often formally acknowledged by the acronym BTEX.

BenzeneTolueneEthylbenzenep-Xylenem-Xyleneo-Xylene
Molecular formulaC6H6C7H8C8H10C8H10C8H10C8H10
Molecular mass78.1292.15106.17106.17106.17106.17
Boiling point, °C80.1110.6136.2138.4139.1144.4
Melting point, °C5.5–95.0–95.013.3–47.9–25.2

t0050

The three isomeric xylenes (1,2-CH3C6H4CH3, 1,3-CH3C6H4CH3, and 1,4-CH3C6H4CH3) and another isomer, ethylbenzene (C6H5C2H5) can be separated only with difficulty, but numerous separation methods have been worked out. The small letters o-, m-, and p- (standing for ortho-, meta-, and para-) preceding the name xylene are used to identify the three different isomers that vary in the ways the two methyl groups displace the hydrogen atoms of benzene. Ortho-xylene is used mostly to produce phthalic anhydride, an important intermediate that leads principally to various coatings and plastics. The least valued of the isomers is meta-xylene, but it has uses in the manufacture of coatings and plastics. Para-xylene leads to polyesters, which reach the ultimate consumer as polyester fibers under various trademarked names.

BTX chemistry focuses on the chemistry of benzene, toluene, and the xylene isomers (Fig. 3.1). Styrene, discussed under C-2 chemistry, is one of the main industrial chemicals made from benzene. Most benzene is alkylated with ethylene to form ethylbenzene, which is dehydrogenated to styrene.

C6H6C6H5C2H5C6H5CHCH2

si15_e

Benzene is an important intermediate in the manufacture of industrial chemicals (Table 3.7) and the products from benzene are frequently feedstocks for the synthesis of additional organic chemicals. Chemically benzene, which forms the basis of the aromatics and considered a human carcinogen by the US EPA, is a closed, six-sided ring structure of carbon atoms with a hydrogen atom at each corner of the hexagonal structure. Thus, a benzene molecule is made up of six carbon (C) atoms and six hydrogen (H) atoms and has the chemical formula C6H6. Benzene is the simplest and most stable aromatics and is a single-ring system that contains conjugated double bonds (Chapter 2):

Table 3.7

Industrial Uses of Benzene

ProcessTarget ProductProcess ConditionsReaction ComponentsOther Characteristics
Pressure (MPa)Temperature (°C)Catalyst
OxidationPhenol0.690–100Cumene, oxygenMost important phenol synthesis
Maleic anhydride0.1–0.2350–400Vanadium oxideButane oxygen
Styrene0.1580–590Iron oxideEthylene benzene
Addition
AlkylationEthylbenzene0.2–0.4125–140Aluminum chlorideBenzene, ethylenePrecursor to styrene
Ethylbenzene2.0420–430ZeoliteBenzene, ethylenePrecursor to styrene
Cumene0.3–1.0250–350Phosphoric acid silicateBenzene, propylene
2.6-Xylenol0.1–0.2300–400Aluminum oxidePhenol, methanol
HydrogenationCyclohexanone0.1140–170PalladiumPhenol, hydrogen
Cyclohexanol1.0–2.0120–200Nickel/silicon oxide and aluminum oxidePhenol, hydrogen
Cyclohexane2.0–5.0150–200NickelBenzene, hydrogen
Aniline0.18270CopperNitrobenzene, hydrogen
NitrationNitrobenzene0.160Benzene, sulfuric acid, nitric acid
SulfonationSurfactants0.140–50Alkylbenezenes/sulfur trioxide
ChlorinationChlorobenzene0.130–40Aluminum chloride/Iron chlorideBenzene, chlorine
CondensationBiphenol A0.150–90HClPhenol, acetone

t0040

u03-15-9780128044926
Benzene showing the electronic configurations that stabilize the ring system.

In the early days of the organic chemicals industry, benzene was obtained from the carbonization of coal, which produces combustible gas, coke, combustible gas, as well as a number of by-products, including benzene. Carbonization of coal to produce illuminating gas dates back in England to the very early years of the 19th century and the process is still employed in some countries, but more use is being made of natural gas. The carbonizing process is also used (with some slight modifications) to produce metallurgical coke, indispensable for the manufacture of iron and hence steel.

Toluene differs from benzene in that one of the hydrogen atoms is replaced by a special combination of carbon and hydrogen called a methyl group (CH3).

u03-16-9780128044926
Benzene, toluene and the isomeric xylenes.

Toluene is also used as a solvent—the substance dissolved is usually also an organic compound. The xylene isomers have two methyl groups in different positions in the benzene ring, and thus all aromatics are to some extent interchangeable. In fact, one of the uses for toluene is to produce benzene by removing the methyl group. All of these hydrocarbons are useful as gasoline additives because of their antiknock properties (high octane numbers).

Benzene itself is perhaps the industrial chemical with the most varied uses of all. There are several routes to phenol, itself an important industrial chemical. In transforming benzene to the products obtained from it, other raw materials are required; for example, ethylene for the production of styrene, and sulfuric acid for the production of benzene sulfonic acid.

Cumene (isopropyl benzene) is produced by the Friedel-Crafts alkylation of benzene with propylene using an acid catalyst.

C6H6+CH3CHCH2C6H5CHCH32

si16_e

u03-17-9780128044926

Although cumene is a high-octane automotive fuel and the high octane number makes it desirable in gasoline, most cumene is easily oxidized to the hydroperoxide, which is readily cleaved in dilute acid to phenol and acetone.

C6H5CHCH32C6H5CO2HCH32C6H5OH+CH32CO

si17_e

Thus, almost all of the cumene produced is used to make phenol (C6H5OH) and acetone [(CH3)2CO] which have a number of important commercial uses, but they also have an important use together. Phenol and acetone can be condensed to form bisphenol A, which is used in the production of polycarbonate and epoxy resins.

u03-18-9780128044926

Acetone [(CH3)2Cglyph_dbndO], a highly volatile and flammable organic chemical, is irritating to the eyes, nose, and throat. Symptoms of exposure to large quantities of acetone may include headache, unsteadiness, confusion, lassitude, drowsiness, vomiting, and respiratory depression. Reactions of acetone in the lower atmosphere contribute to the formation of ground-level ozone. Ozone (a major component of urban smog) can affect the respiratory system, especially in sensitive individuals such as asthmatics or allergy sufferers. If released into water, acetone will be degraded by microorganisms or will evaporate into the atmosphere, although degradation by microorganisms will be the primary removal mechanism. Once acetone reaches the troposphere (the layer known as the lower atmosphere), it will react with other gases, contributing to the formation of ground-level ozone (O3) and other air pollutants.

Benzene is also an important starting material for the manufacture of cyclohexane (C6H12). The process involves the hydrogenation of benzene (over a nickel or platinum catalyst). However, most of the cyclohexane is converted to adipic acid [(CH2)4(COOH)2, hexanedioic acid, hexane-1,6-dicarboxylic acid, hexane-1,6-dioic acid]:

u03-19-9780128044926

The process is an oxidation process in which the intermediate chemicals cyclohexanol and cyclohexanone are oxidized to the adipic acid:

HOC6H11Cyclohexanol+OC6H10O4Adipicacid+HNO2+H2O

si18_e

OC6H10Cyclohexanone+OC6H10Adipicacid+H2O

si19_e

In addition, adipic acid can be reacted with 1,6-hexamethylenediamine (H2NCH2CH2CH2CH2CH2CH2NH2) to produce nylon-6,6, a very strong synthetic fiber.

nHOOCCH24COOHAdipicacid+nH2NCH26NH2HexamethylenediamineOCCH24CONHCH26NHnNylon-6,6+2nH2O

si20_e

Most carpets are made of nylon, as are many silk-like garments, some kinds of rope, and many injection-molded articles. Caprolactam (C6H11NO) is also used to make nylon. Nylon-6,6 is made by direct polymerization of caprolactam, often obtained by reaction of cyclohexanone with hydroxylamine, followed by rearrangement of the oxime. Although nylon-6,6 is the dominant nylon produced in the United States, nylon-6 is the leading nylon product in Europe.

Aniline (C6H5NH2) is made by nitration of benzene to nitrobenzene followed by hydrogenation of the nitrobenzene using a copper-based catalyst in which the copper is suspended in silica (Cu/SiO2) to produce aniline:

C6H6C6H5NO2Nitrobenzene

si21_e

C6H5NO2C6H5NH2Aniline

si22_e

The major use of aniline is in making diisocyanates (RNglyph_dbndCglyph_dbndO), which are used in producing polyurethane materials (such as home insulation).

Alkylbenzene sulfonates (RC6H5SO3Na) are important surfactant compounds used in laundry detergents. In the manufacturing process, alkylbenzenes (made by the Friedel-Crafts alkylation of benzene using linear olefin molecules that have up to 12 carbon atoms) are sulfonated, and the sulfonic acids are then neutralized with sodium hydroxide (NaOH).

Toluene diisocyanate (TDI) is polymerized with diols to produce polyurethanes, which are used to make flexible foam for furniture cushions, mattresses, and carpet pads. Trinitrotoluene (TNT) is made via a stepwise nitration of toluene in the 2, 4, and 6 positions—TNT is a high explosive and missile propellant. Phthalic anhydride is made by air oxidation of ortho-xylene. Approximately 50% w/w of the phthalic anhydride produced is used to make plasticizers, especially the compound dioctyl phthalate [also called bis(2-ethylhexyl) phthalate, di-2-ethylhexyl phthalate, diethylhexyl phthalate, DEHP] for softening PVC plastic. Phthalic anhydride is also used to make unsaturated polyester resins and alkyd paints.

u03-20-9780128044926
Phthalic anhydride [bis(2-ethylhexyl) phthalate, di-2-ethylhexyl phthalate, diethylhexyl phthalate, DEHP]
u03-21-9780128044926
Dioctyl phthalate

3.6 Other Chemical Reactions

In addition to the C-1, C-2, C-3, and C-4 chemistries, there is a wide variety of chemical reactions that are used to produce organic chemicals, some of which are specific to one or two products, whilst others (e.g., oxidation, halogenation, hydrogenation) are used widely in many processes. For this reason, the majority of emissions from the industrial production of organic chemicals originate from a relatively few, but commonly used, unit processes. These reactions are presented here (alphabetically rather than by order of inferred importance) for information purposes.

3.6.1 Alkylation

Alkylation is the introduction of an alkyl group into an organic compound by substitution or addition. There are six types of alkylation reaction: (1) substitution for hydrogen bound to carbon, such as ethylbenzene (C6H5C2H5) from benzene (C6H6) and ethylene (CH2glyph_dbndCH2), (2) substitution for hydrogen attached to nitrogen, (3) substitution for hydrogen in a hydroxyl group of an alcohol or phenol, and (4) addition to a tertiary amine to form a quaternary ammonium compound. Acylation reactions that fall outside of the context of this text—such as the addition to a metal to form a carbon-metal bond as well as additions to sulfur or silicon—are not included in this list. The greatest use of the alkylation process is in refineries for the production of alkylates that are used as a blending stock to produce gasoline (Speight, 2014a,b). Other major alkylation products include ethylbenzene (C6H5CH2CH3), cumene [C6H5CH(CH3)2], and linear alkylbenzene derivatives [C6H5(CH2)nCH3, where n is typically 3 or greater].

Alkylation is commonly carried out in liquid phase at temperatures higher than 200°C (390°F) at above atmospheric pressures. Sometimes vapor phase alkylation is more effective. Alkylation agents are usually olefins, alcohols, alkyl sulfates, or alkyl halides and the typical catalysts are hydrofluoric acid (HF), sulfuric acid (H2SO4), or phosphoric acid (H3PO4, which is also known as orthophosphoric acid or phosphoric(V) acid). Higher process temperatures cause the expected lowering of product specificity and increased by-product formation. Some more recent alkylation processes (such as used for the production of ethylbenzene and cumene) use zeolite catalysts as they can be more efficient and may have lower emissions. Lewis acids, such as aluminum chloride (AlCl3) or boron trifluoride (BF3), may also be used as catalysts.

There are environmental issues, related to the emission of volatile organic compounds (VOCs), which arise during alkylation processes and which include although based on data for the production of ethylbenzene, cumene, and linear alkylbenzene, the emission of VOCs from alkylation reactions tend to be low compared to the emission of VOCs from other refinery processes. However, the by-products and waste disposal of alkyl halides and sulfate derivatives can be problematic.

3.6.2 Ammonolysis

Ammonolysis is the process of forming amines using, as amminating agents, ammonia or primary and secondary amines. These reactions may also include hydroammonolysis in which amines are formed directly from carbonyl compounds using an ammonia-hydrogen mixture and a hydrogenation catalyst. The four main ammonolysis reaction types are: (1) double decomposition in which the ammonia (NH3) is split into the amino function (NH2) which becomes part of the amine, and a hydrogen atom which reacts with a radical that is being substituted, (2) dehydration, in which the ammonia reacts to produce water and amines, (3) simple addition, in which both fragments of the ammonia molecule (glyph_sbndNH2 and glyph_sbndH) become part of the new amine, and (4) multiple activity, in which ammonia reacts with the amine products to form secondary amine derivatives and tertiary amine derivatives. Nevertheless, typically, the major products of ammonolysis are carbamic acid (H2NCOOH), ethanolamine derivatives (HOCH2CH2NHR), and alkylamine (RNH2) derivatives.

In the case of the production of ethanolamine, the emission of VOCs arising from the reactors is minimal although there tends to be waste gases associated with any distillation. Any off-gas containing ammonia or amine derivatives is washed or incinerated in order to avoid odor problems and any hydrogen cyanide that is produced during the production of acrylonitrile (CH2glyph_dbndCHCN) can may be recovered.

In the case of water contamination, unreacted ammonia can be recovered from alkaline effluents by steam stripping and recycled back to the process. Ammonia remaining in the effluent can be neutralized with sulfuric acid to produce a precipitate of ammonium sulfate [(NH4)2SO4] that can be separated for use as fertilizer or biologically treated. In addition, wastewater containing impurities such as methanol and amine derivatives can be disposed of either by incineration or by biological treatment. Solid waste materials from base of the stripping unit are incinerated.

3.6.3 Ammoxidation

In the organic chemicals industry, ammoxidation is a process for the production of nitrile derivatives involving the use of ammonia and oxygen—the usual feedstock is an alkene. The process is an important application in the production of acrylonitrile (CH2glyph_dbndCHCN) and process involves the gas phase oxidation of olefins, such as propylene with ammonia in the presence of oxygen and vanadium or molybdenum based catalysts:

2CH3CHCH2+2NH3+3O22CH2CHCN+6H2O

si23_e

Acetonitrile (CH3Cglyph_tbndN) is a by-product of this process. This colorless liquid is the simplest organic nitrile and is used as a polar aprotic solvent in organic synthesis and in the purification of butadiene. Acetonitrile is also a common two-carbon building block in the synthesis of organic chemicals—the most important applications are in the production of acrylic fibers, thermoplastics, and adiponitrile (Nglyph_tbndCCH2CH2CH2CH2Cglyph_tbndN, a viscous, colorless liquid, an important precursor to the polymer Nylon-6,6), as well as specialty polymers.

3.6.4 Carbonylation

Carbonylation (carboxylation) is the combination of an organic compound with carbon monoxide and carbonylation refers to reactions that introduce carbon monoxide into organic and inorganic compounds. The carbonylation reaction is used to make aldehydes and alcohols containing one additional carbon atom—the major product involves the introduction of the carbonyl group (> Cglyph_dbndO) into the starting materials to produce aldehydes (glyph_sbndCHO), carboxylic acids (glyph_sbndCO2H), and esters (glyph_sbndCO2R). In the laboratory and in industry, carbonylation reactions are the basis of two main types of reactions, hydroformylation and Reppe Chemistry.

The hydroformylation reaction involves the addition of both carbon monoxide and hydrogen to unsaturated organic compounds, usually alkene derivatives to produce aldehyde derivatives:

RCHCH2+H2+CORCH2CH2CHO

si24_e

The reaction requires the presence of metal catalysts to bond the carbon monoxide and the hydrogen to the olefin. The hydroformylation (“oxo” process) is a variant where olefins are reacted with carbon monoxide and hydrogen (“synthesis gas”) in the presence of a cobalt or rhodium catalyst (such as in the production of n-butyraldehyde from propylene):

CH3CHCH2+H2+COCH3CH2CH2CHO

si25_e

On the other hand, the Reppe reaction involves the addition of carbon monoxide and an acidic hydrogen donor with the organic substrate. Large-scale applications of this type of carbonylation are the conversion of methanol to acetic acid:

CH3OH+COCH3CO2H

si26_e

In a related hydrocarboxylation, alkene derivatives are converted to carboxylic acid derivative in the presence of metal catalysts:

RCHCH2+H2O+CORCH2CH2CO2H

si27_e

For example, as part of the industrial synthesis of ibuprofen, a benzyl alcohol derivative is converted to the corresponding carboxylic acids by way of palladium-catalyzed carbonylation reaction:

ArCHCH3OH+COArCHCH3CO2H

si28_e

Also, propanoic acid is mainly produced by the hydrocarboxylation of ethylene using nickel carbonyl [nickel tetracarbonyl, Ni(CO)4] as the catalyst:

CH2CH2+H2O+COCH3CH2CO2H

si29_e

Hydroesterification is similar to hydrocarboxylation, but uses alcohol derivatives instead of water. Other related industrially oriented reactions include the Koch reaction, which involves the addition of carbon monoxide to unsaturated compounds in the presence of a catalyst, such as in the production of the important intermediate, glycolic acid:

HCHOFormaldehyde+CO+H2OHOCH2CO2HGlycolicacid

si30_e

Alkyl, benzyl, vinyl, aryl, and allyl halides can also be carbonylated in the presence carbon monoxide and suitable catalysts such as manganese (Mn), iron (Ni), or nickel (Ni).

In terms of environmental issues of the carbonylation processes, the processes typically generate vent streams containing VOCs in addition to carbon dioxide as well as other nonVOCs. Residual gas is recovered and used as fuel or flared. In addition, heavy metals (from the catalyst) must be removed from wastewater prior to biological treatment. The solid waste is typically the spent catalysts but caution is advised when planning the disposal process because of the potential for adsorbed material, which might be leached from the solid during periods of heavy rain, melting snow, or acid rain.

Waste: Spent catalysts.

3.6.5 Condensation

A condensation reaction, is a chemical reaction in which two reactants (each reactant contains a functional group) combine to form a higher molecular weight product together with the loss of a lower molecular weight species, such as water, hydrogen chloride, methanal, or acetic acid—typically, the most common lower molecular weight species is water.

A common type of industrial condensation reaction is a condensation polymerization reaction in which a series of condensation steps takes place whereby monomers or monomer chains add to each other to form longer chains.

HOOCR1COOHDi-acid+H2NR2NH2DiamineCOR1CONHR2NHPolyamiden+nH2O

si31_e

In this reaction, R1 and R2 are alkyl or aryl moieties and the glyph_sbndCONHglyph_sbnd moiety is the amid bond. The reaction is also known as step-growth polymerization and is used in processes such as the synthesis of polyesters or nylons. This reaction may be either (1) homopolymerization such as in the use of a single monomer with two different end groups that condense or (2) copolymerization of two comonomers with the necessary functional groups on each of the monomers (as shown in the equation above). Small molecules are usually liberated in these condensation steps, unlike polyaddition reactions such as in the polymerization of ethylene which is an addition reaction and there is no elimination of lower molecular weight by-products:

nCH2CH2CH2CH2n

si32_e

Environmental issues that arise during the use of condensation processes are usually limited to reactor emissions that are generally small and can be mitigated in a combustion unit. In addition, waste water volumes are generally low and the effluents mainly consist of reaction water if recycling after phase separation is not possible. The effluent is composed of high-boiling components (condensation products/by-products) that often show moderate or poor biodegradability, and low-boiling components that typically are more susceptible to biodegradation.

3.6.6 Dealkylation

Dealkylation is a chemical process through which alkyl groups are removed from a given compound such as the dealkylation of toluene to produce benzene or the conversion of 1,2,4-trimethylbenzene to xylene:

C6H5CH3+H2C6H6+CH41,2,4-C6H3CH33+H2C6H4CH32+CH4

si33_e

When hydrogen is used (above equation) the process is known as hydrodealkylation. The hydrodealkylation process typically requires a high temperature and a high pressure as well as the presence of catalyst—the catalyst is predominantly a transition metal-containing catalyst using metal derivatives of chromium or molybdenum.

The dealkylation process is a common process in the organic chemicals industry, especially in industries such as crude oil refining and the pharmaceuticals industry. In fact, in the pharmaceuticals industry dealkylation can lead to activation of certain compounds and can also promote better absorption and efficacy of the pharmaceutical. The dealkylation process is also frequently employed by manufacturers of fertilizers and pesticides.

Many dealkylation processes use oxidative dealkylation (also known as O-dealkylation) in which an oxide is used to assist in removal of the alkyl group from the organic substrate through a reduction-oxidation reaction (redox reaction).

3.6.7 Dehydration

The dehydration of organic chemicals is a decomposition reaction in which a new compound is formed by the expulsion of water. The dehydration process is a subprocess of the condensation process that requires a catalyst. Examples of common dehydrating reactions include:

ReactionsEquations
Conversion of alcohols to ethers2ROH → ROR + H2O
Conversion of alcohols to alkeneRCH2CHOH-R → RCHglyph_dbndCHR + H2O
Conversion of carboxylic acids to acid anhydrides2RCOOH → (RCO)2O + H2O
Conversion of amides to nitrilesRCONH2 → RCN + H2O

The reverse of a dehydration reaction is a hydration reaction in which water is added to a substrate such as an olefin.

Dehydration:

C2H5OHCH2CH2+H2O

si34_e

Hydration:

CH2CH2+H2OC2H5OH

si35_e

Typical dehydrating agents used in organic synthesis include concentrated sulfuric acid (H2SO4), concentrated phosphoric acid (H3PO4), and aluminum oxide (Al2O3). In the related condensation reaction, water is released from two different reactants.

3.6.8 Dehydrogenation

Dehydrogenation is the process by which hydrogen is removed from an organic compound to form a new chemical (e.g., to convert saturated into unsaturated compounds). It is used to produce aldehydes and ketones by the dehydrogenation of alcohols. Important products include acetone, cyclohexanone, methyl ethyl ketone, and styrene.

Dehydrogenation is most important in the refinery cracking process, where saturated hydrocarbons are converted into olefins (Speight, 2014a,b, 2016). The process is applied to appropriate hydrocarbon feedstocks (e.g., naphtha) in order to produce the very large volumes of ethylene, propylene, butene derivatives, and butadiene derivatives that are required as feedstocks for the organic chemicals industry.

Cracking (thermal decomposition) of organic compounds may be achieved by thermal (noncatalytic) processes or by catalytic processes and provides a process to convert higher boiling fractions into saturated, nonlinear paraffinic compounds, naphthenes, and aromatics. The concentration of olefin derivatives in the product stream is very low, so this method is more useful for the preparation of blending stocks (such as naphtha and kerosene) for the production of fuels (gasoline and diesel fuel, respectively). Olefin derivatives are more widely produced by the steam cracking of petroleum fractions—in the process a hydrocarbon stream is heated, mixed with steam and, depending on the feedstock, further heated to a cracking-temperature on the order of 600–650°C (1110–1200°F). The conversion of saturated hydrocarbon streams to unsaturated compounds is highly endothermic, and so the process requires a high energy input. High-temperature cracking is also used to produce pyrolysis gasoline from naphtha, gas oil, or high-boiling refinery streams.

Environmental issues that arise from the of dehydrogenation processes include the potential for hydrogen-rich vent streams that are produced as a result of the process but which can be employed as a hydrogen feedstock for other processes or as a refinery fuel. Any volatile hydrocarbons that occur in purge and vent gases will require collection and treatment and can be combined with beneficial energy production. On the other hand, quench water, dilution steam, decoking water, and flare water discharges will require treatment and wastewater streams with a high content of pollutants will require pretreatment prior to acceptance in a biological degradation plant.

3.6.9 Esterification

Esterification typically involves the formation of esters from an organic acid and an alcohol. Esters often have a characteristic pleasant, fruity odor which makes them appropriates for extensive use in the fragrance and flavor industry.

The most common method of esterification is the reaction of a concentrated alcohol and a concentrated carboxylic acid with the elimination of water:

R1CO2HAcid+R2OHAlcoholR1CO2R2Ester+H2O

si36_e

In this equation, R1 and R2 are alkyl moieties. Only strong carboxylic acids react sufficiently quickly without a catalyst, so a strong mineral acid (such as sulfuric acid or hydrogen chloride) must usually be added to aid the reaction. Acid anhydrides are also used, e.g., in dialkyl phthalate production. The sulfonic acid group can be bound chemically to a polymeric material and so cation exchangers, such as sulfonated polystyrene, enable esterification under mild conditions.

The equilibrium of the reaction can be shifted to the ester by increasing the concentration of one of the reactants, usually the alcohol. In production scale esterification the reaction mixture is refluxed until all the condensation water is formed, and the water or the ester product is continuously removed from the equilibrium by distillation. The main products from esterification reactions are dimethyl terephthalate, ethyl acrylate, methyl acrylate, and ethyl acetate which have considerable economic importance in many applications (such as for fibers, films, adhesives and plastics). Some volatile esters are used as aromatic materials in perfumes, cosmetics, and foods.

Also, the reaction of alcohol derivatives with carboxylic acid derivatives is not the only process for producing ester derivatives. Alcohols react with acyl chlorides (acid chlorides) and acid anhydrides to produce esters:

R1COCl+R2OHR1CO2R2+HClR1CO2O+R2OHR1CO2R2+R1CO2H

si37_e

Again, in these equations R1 and R2 are alkyl moieties. The reactions are irreversible, simplifying and driving the process to completion. Since acyl chlorides and acid anhydrides also react with water, anhydrous conditions are preferred. The analogous acylation of amine derivatives to yield amide derivatives are less sensitive because amines are stronger nucleophiles and react more rapidly than does water.

Finally, ethylene, acetic acid, and oxygen react (in the presence of palladium-based catalysts) to yield vinyl acetate:

2CH2CH2+2CH3CO2H+O22C2H3O2CCH3+2H2O

si38_e

Direct routes (such as the alcohol-acid reaction) to this same ester are not possible because vinyl alcohol (CH2glyph_dbndCHOH) is unstable and has the propensity under normal conditions (ambient temperature and ambient pressure) to convert to acetaldehyde (CH3CHO).

Environmental issues related to the esterification process relate to solvent vapor which can be collected and treated (such as by incineration, adsorption). The generation of aqueous effluents is not extensive since water is the only by-product of the esterification reaction. In addition, the choice of solid polymer based ion exchange resins for wastewater treatment avoids the need for extensive treatment facilities. Most esters possess low toxicity because they are easily hydrolyzed on contact with water or moist air, and so the properties of the acid and alcohol components are more important. Furthermore, waste streams can be reduced by recovering (and reusing) any organic solvents, water, and alcohol components. Any wastes from waste water treatment can be incinerated (high-boiling wastes) or recovered by distillation for reuse (low-boiling wastes).

3.6.10 Halogenation

Generally, halogenation is the reaction of a halogen with an alkane in which the introduction of halogen atoms occurs into the organic molecule by an addition reaction or by a substitution reaction. In organic synthesis this may involve the addition of molecular halogens: chlorine, bromine, iodine, or fluorine (Cl2, Br2, I2, or F2) or hydrohalogenation using: hydrogen chloride, hydrogen bromide, hydrogen iodide, or hydrogen fluoride (HCl, HBr, HI, or HF) to carbon-carbon double bonds. Substitution reactions involve replacing hydrogen atoms in olefin derivatives, paraffin derivatives, or aromatic derivatives with halogen atoms. Chlorination is the most important industrial halogenation reaction and chlorinated organic products include chlorinated aromatic derivatives, chlorinated methane derivatives, and chlorinated ethane derivatives but caution is advised since toxicity issues will demand additional control measures. Fluorination is used almost exclusively in the manufacture of fluorocarbons.

Several pathways exist for the halogenation of organic compounds, including free radical halogenation, electrophilic halogenation, and the halogen addition reaction. For example, saturated hydrocarbon derivatives (alkanes) typically do not add halogens but undergo free radical halogenation which involves the substitution of a hydrogen atom (or hydrogen atoms) by a halogen atom (or halogen atoms). The chemistry of the halogenation of alkane derivatives is usually determined by the relative weakness of the available carbon-hydrogen (Cglyph_sbndH) bonds. Free radical halogenation is used for the industrial production of chlorinated methane derivatives:

CH4+Cl2CH3Cl+HCl

si39_e

Rearrangement often accompanies such free radical reactions. On the other hand, unsaturated compounds, especially alkene derivatives and alkyne derivatives add halogens across the unsaturated bond:

R1CHCHR2+X2R1CHXCHXR2

si40_e

However, aromatic compounds are subject to electrophilic halogenation but addition to the ring system can occur under extreme conations:

RC6H5+X2RC6H4X+HX

si41_e

The ease of the reaction is influenced by the halogen—fluorine and chlorine are more electrophilic and, as a result, are more aggressive halogenating agents. On the other hand, bromine is a weaker halogenating agent than both fluorine and chlorine, while iodine is least reactive halogenating agent of the halogens. Accordingly, the ease of hydrogenolysis (removal of the halogen with hydrogen as HX) follows the reverse trend: iodine is most easily removed from organic compounds and organo fluorine compounds are most stable organo-halogen compounds.

Environmental issues of halogenation processes involve the treatment of waste gases which requires a distinction between acidic streams, reaction gases, and neutral waste streams. Air streams from tanks, distillation columns, and process vents can be collected and treated using such techniques as low temperature condensation or incineration. However, the treatment of acid gas streams is more complex because any equipment in contact with acid gases and water must be constructed from acid-resistant materials or internally coated to prevent corrosion (Speight, 2014c). The halogen content of the waste gas may represent a valuable raw material and pollution control techniques offer an opportunity for its recovery and reuse (either as hydrogen-halogen or aqueous solutions). The techniques may include: (1) product recovery (by vapor stripping of liquid streams followed by recycling to the process), (2) scrubbing the acid gas with an easily halogenated compound preferably a raw material used in the process, (3) absorbing the acid gas in water to give aqueous acid which is often followed by caustic scrubbing for environmental protection, (4) washing out organic constituents with organic solvents, and (5) condensing out organic by-products for use as feedstock in another process (Mokhatab et al., 2006; Speight, 2007, 2014a).

Environmental issues also arise with wastewater streams because the biological degradability (biodegradability) of halogenated hydrocarbons (especially aromatic derivatives) decreases as the halogen content increases. Only chlorinated hydrocarbon derivatives with a low degree of chlorination are degradable in biological waste water treatment plants but only if the concentration of the chlorinated hydrocarbon derivatives does not exceed certain levels. Prior to biological treatment, wastewater containing chlorinated organic compounds usually requires preliminary purification by stripping, extraction, and adsorption (using activated carbon or suitable polymeric resins). Wastewater contamination can be substantially reduced by avoiding the water quenching of reaction gases to separate hydrogen chloride (for example in the production of chlorinated ethane derivatives and chlorinated ethylene derivatives).

Finally, solid waste materials as a result of the halogenation process may arise from sources such as reactor residues or spent catalyst. Incineration is a common method for destruction of the organic components of the solid wastes but considerable attention must be paid to incineration conditions in order to avoid the formation of dioxins. If incineration is used, there is the need for an efficient flue gas scrubbing operation.

3.6.11 Hydrogenation

Catalytic hydrogenation refers to the addition of hydrogen to an organic molecule in the presence of a catalyst. The process can involve direct addition of hydrogen to the double bond of an unsaturated molecule; amine formation by the replacement of oxygen in nitrogen-containing compounds; and alcohol production by addition to aldehydes and ketones. These reactions are used to readily reduce many functional groups; often under mild conditions and with high selectivity.

ReactantProduct
Alkene: R2Cglyph_dbndCR′2Alkane: R2CHCHR′2
Alkyne: RCCRAlkene: cis-RHCglyph_dbndCHR′
Aldehyde: RCHOPrimary alcohol: RCH2OH
Ketone: R2COSecondary alcohol: R2CHOH
Ester: RCO2R′Mixed alcohols: RCH2OH + R′OH
Imine: RR′CNR″Amine: RR′CHNHR″
Amide: RC(O)NR′2Amine: RCH2NR′2
Nitrile: RCNPrimary amine: RCH2NH2
Nitro: RNO2Amine: RNH2

Hydrogenation is an exothermic reaction and the equilibrium usually lies far towards the hydrogenated product under most operating temperatures. It is used to produce a wide variety of chemicals such as cyclohexane, aniline, n-butyl alcohol, hexamethylene diamine, as well as ethyl hexanol, and important isocyanate derivatives such as TDI and methylene diphenyl isocyanate both of which are used to produce urethane derivatives and thence urethane polymers.

Hydrogenation catalysts may be heterogeneous or homogeneous—heterogeneous catalysts are solids and form a distinct phase in the gases or liquids. Many metals and metal oxides have general hydrogenation activity—nickel, copper, cobalt, chromium, zinc, iron, and the platinum group are among the elements most frequently used as commercial hydrogenation catalysts.

Generally, the emission of VOCs for hydrogenation processes is relatively low since hydrogen-rich vent streams are typically sent to combustion units. The main issues with hydrogen are likely to arise from sulfur impurities in the process feedstocks or from the dust and ash by-products of the hydrogen production itself. Small quantities of sulfur compounds (such as sulfur dioxide, SO2, and hydrogen sulfide, H2S) can be absorbed in dilute caustic solutions or adsorbed on activated charcoal as part of a gas cleaning operation while larger quantities would probably have to be converted to liquid sulfur or to solid sulfur (Mokhatab et al., 2006; Speight, 2007).

The hydrogenation of oxygenated compounds may generate water, which ends up as waste water, but volume of wastewater produced from hydrogenation reactions is not excessive. Moreover, the products often show good biodegradability and low toxicity whereas aniline compounds (from a hydrodenitrogenation process) will need disposal measures that are additional to the bio treatment technologies. In addition, the spent catalysts may be sent to disposal or may be treated for reclamation of any precious metals.

3.6.12 Hydrolysis

Hydrolysis involves the reaction of an organic chemical with water to form two or more new substances and usually means the cleavage of chemical bonds by the addition of water. In fact, Hydrolysis can be the reverse of a condensation reaction in which two molecules join together into a larger one and eject a water molecule. Thus hydrolysis adds water to break down, whereas condensation builds up by removing water. Hydration is the process variant where water reacts with a compound without causing its decomposition. These routes are used in the manufacture of alcohols (e.g., ethanol), glycols (e.g., ethylene glycol, propylene glycol), and propylene oxide.

Acid-base-catalyzed hydrolysis reactions and processes are very common—one example is the hydrolysis of ester derivatives or amide derivatives. The hydrolysis reaction occurs when the nucleophilic reactant (a nucleus-seeking agent, e.g., water or hydroxyl ion) attacks the carbon of the carbonyl group of the ester or the amide using an aqueous base medium since hydroxyl ions are better nucleophiles than polar molecules such as water. In acidic medium, the carbonyl group becomes protonated, and this leads to a much easier nucleophilic attack. The products for both hydrolyses are carboxylic acid derivatives. The oldest commercially practiced example of ester hydrolysis is the saponification reaction which results in the formation of soap and involves the hydrolysis of a triglyceride (fat) with an aqueous base such as sodium hydroxide (NaOH). During the process, glycerol (CH2OHCHOHCH2OH) is formed:

u03-22-9780128044926

The carboxylic acids react with the base, converting them to salts.

In the hydrolysis process, there are generally low-to-no emission of VOCs emanating from the reactor and, in most cases, the products of the hydrolysis process and the hydration process are biodegradable.

3.6.13 Nitration

The nitration process is a general class of chemical process for the introduction of a nitro group (glyph_sbndNO2) into an organic chemical compound. The term is also applied (somewhat incorrectly) to the different process of forming nitrate esters between alcohol derivatives and nitric acid, as occurs in the synthesis of nitroglycerin. The difference between the resulting structure of nitro compounds and nitrates is that the nitrogen atom in nitro compounds is directly bonded to a nonoxygen atom, typically carbon or another nitrogen atom, whereas in nitrate esters, also called organic nitrates, the nitrogen is bonded to an oxygen atom that in turn usually is bonded to a carbon atom. There are many major industrial applications of nitration in the strict sense; the most important by volume are for the production of nitroaromatic compounds such as nitrobenzene. Nitration reactions are notably used for the production of explosives, for example the conversion of toluene to TNT (2,4,6-trinitrotoluene).

u03-23-9780128044926

However, explosives aside, the nitro compounds are of wide importance as chemical intermediates and precursors.

Thus, nitration involves the replacement of a hydrogen atom (in an organic compound) with one or more nitro groups (glyph_sbndNO2). By-products may be unavoidable due to the high reaction temperatures and the highly oxidizing environment, although many nitration reactions are carried out at low temperature for safety reasons. The nitration reaction can be carried out with aliphatic compounds (to produce nitroparaffin derivatives) but the nitration of aromatics is more commercially important (to produce explosives and propellants such as nitrobenzene and nitrotoluene derivatives). This is effected with nitric acid (HNO3) or, in the case of aromatic nitration reactions, a mixture of nitric and sulfuric acids. Nitration is used in the first step of TDI production.

Environmental issues of nitration processes (excluding the more obvious potential explosive properties) relate to the occurrence of acid vapors (largely nitric or sulfuric acid) from the reaction and quenching as well as any unreacted nitrating agent arising from the use of an excess of the agent to carry the nitration reaction to completion. There is also the potential for the emission of VOCs as well as other gas streams that contain the various oxides of nitrogen. In terms of water pollutants, the nitration of aromatic feedstocks may produce large quantities of waste mixed acid that requires neutralization and disposal, or recovery (e.g., by distillation) and reuse. The products and by-products of the nitration process often are slow to biodegrade (if they are at all biodegradable) and toxic, so additional treatment of the waste products (such as extraction or incineration of aqueous wastes) may be required.

3.6.14 Oxidation

The term oxidation includes many different processes, but in general it describes the addition of one or more oxygen atoms to a compound. Catalytic oxidation process are processes that utilize catalysts to enhance the oxidation reaction—typical oxidation catalysts are metal oxides and metal carboxylates. The catalysis of the oxidation process occurs by the use of both heterogeneous catalysis and homogeneous catalysis (Table 3.8). In the heterogeneous processes, gaseous substrate and oxygen (or air) are passed over solid catalysts. Typical catalysts are platinum, redox-active oxides of iron, vanadium, and molybdenum. In many cases, catalysts are modified with the suitable choice (from an extensive list) of additives or promoters that enhance the reaction rate or the product selectivity.

Table 3.8

Examples of Industrial Oxidation Processes

SubstrateProcessCatalystProductApplication
ButaneMaleic anhydride processVanadium phosphates (heterogeneous)Maleic anhydridePlastics, alkyd resins
CyclohexaneK-A processCo and Mn salts(homogeneous)Cyclohexanol, cyclohexanoneNylon precursor
EthyleneEpoxidationMixed Ag oxides(heterogeneous)Ethylene oxideBasic chemicals, surfactants
EthyleneOMEGA processEthylene glycol
EthyleneWacker processPd and Cu salts (homogeneous)AcetaldehydeBasic chemicals
MethanolFormox processFe-Mo-oxides (heterogeneous)FormaldehydeBasic chemicals, alkyd resins
PropyleneAllylic oxidationMo-oxides (heterogeneous)Acrylic acidPlastic precursor
Propylene, ammoniaSOHIO processBi-Mo-oxides (heterogeneous)AcrylonitrilePlastic precursor
p-XyleneTerephthalic acid synthesisMn and Co salts (homogeneous)Terephthalic acidPlastic precursor

t0045

The important homogeneous catalysts for the oxidation of organic compounds are the carboxylic acid salts of cobalt, iron, and manganese. To confer good solubility in the organic solvent, these catalysts are often derived from naphthenic acids and ethylhexanoic acid which are highly lipophilic. These catalysts initiate radical chain reactions that produce organic radicals that combine with oxygen to give hydroperoxide intermediates. Generally, the selectivity of oxidation is determined by the bond energy—for example, benzylic carbon hydrogen (C6H5CHglyph_sbndH) bonds are replaced by oxygen faster than aromatic carbon hydrogen (C6H5glyph_sbndH) bonds.

Common applications of the process involve oxidation of organic compounds by the oxygen in air. Such processes are conducted on a large scale for the remediation of pollutants, production of valuable chemicals, and the production of energy. An illustrative catalytic oxidation is the conversion of methanol to the more valuable compound formaldehyde using aerial oxygen:

2CH3OH+O22HCHO+2H2O

si42_e

This conversion is very slow in the absence of catalysts.

Atmospheric oxygen is by far the most important, and the cheapest, oxidizing agent although the inert nitrogen component will dilute products and generate waste gas streams. Other oxidizing agents include nitric acid, sulfuric acid, oleum, hydrogen peroxide, organic peroxides, and pure oxygen. In general terms, organic materials can be oxidized either by heterolytic or homolytic reactions, or by catalytic reactions (where the oxidizing agent is reduced and then reoxidized). Heterogeneous catalysts based on noble metals play a dominant role in industrial scale oxidations and an important example is the silver catalyzed gas phase reaction between ethylene and oxygen to form ethylene oxide. Ethylene is still the only olefin that can be directly oxidized to the corresponding epoxide with high selectivity. Other important industrial oxidation processes are the production of acetic acid, formaldehyde, phenol, acrylic acid, acetone, and adipic acid. Oxidation reactions are exothermic and heat can be reused in the process to generate steam or to preheat other component streams. Fire and explosion risks exist with heterogeneously catalyzed direct oxidation processes (e.g., ethylene oxide process) and reactions involving concentrated hydrogen peroxide or organic peroxides.

In terms of the environmental aspects of the oxidation process, the oxidation of organic compounds produces a number of by-products (including water) and wastes from partial and complete oxidation. In the organic chemical industry, such compounds as aldehydes, ketones, acids, and alcohols are often the final products of partial oxidation of hydrocarbons. Careful control of partial oxidation reactions is usually required to prevent the material from oxidizing to a greater degree than desired as this produces carbon dioxide and many undesirable gaseous, liquid, or semisolid toxic by-products.

In addition, the emissions of volatile organics can arise from losses of unreacted feed, by-products, and products such as aldehydes and acids. Carbon dioxide is an ever-present by-product in the oxidation of organic compounds since it is difficult (if not impossible in some cases) to prevent the complete oxidation of some carbon. Aldehyde derivatives (especially formaldehyde, HCHO) require strict handling to minimize exposure and this limits atmospheric emissions. Acid gases usually require removal from waste streams. Also, to enable biological degradation in a wastewater treatment plant it will be necessary to neutralize any acidic components and to remove/destroy any chlorinated species that may inhibit biological activity.

3.6.15 Oxyacetylation

Acetylation is a reaction that introduces an acetyl functional group (acetoxy group, CH3Cglyph_dbndO) into an organic chemical compound—namely the substitution of the acetyl group for a hydrogen atom—while deacetylation is the removal of an acetyl group from an organic chemical compound. Thus, oxyacetylation involves the addition of oxygen and an acetyl group to an olefin to produce an unsaturated acetate ester. It is used to produce vinyl acetate from ethylene, acetic acid, and oxygen.

3.6.16 Reforming

Reforming is the decomposition (cracking) of hydrocarbon gases or low octane petroleum fractions by heat and pressure. Catalytic reforming is used to convert naphtha (having low octane ratings) into a high-octane liquid product (reformate) which is a premium blending stock for the production of high-octane gasoline. The process converts low-octane linear hydrocarbons (paraffins) into branched alkanes (isoparaffins) and cycloalkane derivatives, which are then partially dehydrogenated to produce high-octane aromatic hydrocarbon derivatives. The dehydrogenation also produces significant amounts of by-product hydrogen, which is used in other refinery processes such as hydrocracking. A side reaction in the reforming process is hydrogenolysis, which produces low-boiling hydrocarbons, such as methane, ethane, propane, and butanes. The nature of the final product is influenced by the source (and composition) of the feedstock. The four major catalytic reforming reactions are:

(1) The dehydrogenation of cycloalkane derivative (naphthenes) to aromatics:

u03-24-9780128044926

(2) The isomerization of normal paraffin derivatives to isoparaffin derivatives:

u03-25-9780128044926

(3) The dehydrogenation and aromatization of paraffin derivatives to aromatic products (known as dehydrocyclization): as exemplified in the conversion of normal heptane to toluene, as shown below:

u03-26-9780128044926

(4) The hydrocarbon of paraffin derivatives into lower molecular weight products

CH3CH2CH2CH2CH2CH2CH3n-HeptaneCH32CHCH2CH3Isopentane+CH3CH3Ethane

si43_e

The hydrocracking of paraffin derivatives is only one of the above four major reforming reactions that consumes hydrogen. The isomerization of normal paraffins does not consume or produce hydrogen but, moreover, both the dehydrogenation of naphthene derivatives and the dehydrocyclization of paraffin derivatives produce hydrogen. In many petroleum refineries, the net hydrogen produced in catalytic reforming supplies a significant part of the hydrogen used elsewhere in the refinery (for example, in hydrodesulfurization processes).

In a variation of the reforming process, the steam reforming is a process for producing hydrogen, carbon monoxide, or other useful products from hydrocarbon feedstocks such as natural gas, which is predominantly methane, hence the alternate name for the process: steam-methane reforming. The conversion is achieved in a reactor (reformer) in which the methane reacts at high temperature with the steam. At high temperatures (700–1100°C, 1290–2010°F) and in the presence of a metal-based catalyst (nickel), steam reacts with methane to yield carbon monoxide and hydrogen:

CH4+H2OCO+3H2

si44_e

Thus:

CnHm+nH2On+m2H2+nCO

si45_e

Additional hydrogen can be recovered by a lower-temperature gas-shift reaction with the carbon dioxide produced, in the presence of a copper-based or iron-based catalyst:

Water-gas shift reaction:

CO+H2OCO2+H2

si46_e

The first reaction is strongly endothermic (consumes heat) while the second reaction is mildly exothermic (produces heat).

3.6.17 Sulfonation

Sulfonation is the process by which a sulfonic acid group (or corresponding salt or sulfonyl halide) is attached to a carbon atom and the process is used to produce detergents (by sulfonating mixed linear alkyl benzenes with sulfur trioxide or oleum). In the process, a hydrogen atom on an aromatic ring is replaced by a sulfonic acid functional group by an electrophilic aromatic substitution reaction:

u03-27-9780128044926

Thus, the general equation for sulfonation of the aromatic ring is:

ArH+H2SO4OleumArSO3H+H2O

si47_e

The most widely used sulfonating agent for linear alkylbenzenes is oleum (fuming sulfuric acid—a solution of sulfur trioxide in sulfuric acid, H2SO4·SO3). Sulfuric acid alone is effective in sulfonating the benzene ring but the acid content of the sulfuric acid (solution) must be in excess of 75% v/v. The excess sulfur trioxide in oleum removes the water of reaction thereby preventing the sulfuric acid from descending below the minimum acid content level for an efficient reaction and promotes higher yields of the desired product. Separating the product sulfonates from the reaction mixture is often difficult—the mother liquor (the solution remaining after the reaction) after product separation raises an environmental issue.

The acid vapor from the reaction and quenching as well as unreacted sulfonating agent arising from the excess use to drive the reaction, pose serious disposal problems. Also, acidic wastewater from the reactor and dilute acidic wash waters (from washing the product on the filter) require neutralization. In addition, the filtrate from the separation stage is contaminated with unreacted raw material and acid. Finally, oleum is an extremely strong oxidizing agent and produces tar by-products that also require disposal.

References

Aiello-Mazzarri C., Coward-Kelly G., Agbogbo F.K., Holtzapple M.T. Conversion of municipal solid wastes into carboxylic acids by anaerobic countercurrent fermentation—effect of using intermediate lime treatment. Appl. Biochem. Biotechnol. 2005;127(2):79–93.

Aiello-Mazzarri C., Agbogbo F.K., Holtzapple M.T. Conversion of municipal solid wastes into carboxylic acids using a mixed culture of mesophilic microorganisms. Bioresour. Technol. 2006;97(1):47–56.

Chadeesingh R. The Fischer-Tropsch process. In: Speight J.G., ed. The Biofuels Handbook. London: The Royal Society of Chemistry; 2011:476–517 (Part 3, Chapter 5).

Chan W.N., Holtzapple M.T. Conversion of municipal solid wastes to carboxylic acids by thermophilic fermentation. Appl. Biochem. Biotechnol. 2003;111(2):93–112.

Chang V.S., Holtzapple M.T. Fundamental factors affecting biomass enzymatic reactivity. Appl. Biochem. Biotechnol. 2000;84(6):5–37.

Chang V.S., Kaar W.E., Burr B., Holtzapple M.T. Simultaneous saccharification and fermentation of lime-treated biomass. Biotechnol. Lett. 2001a;23(16):1327–1333.

Chang V.S., Nagwani M., Kim C.H., Holtzapple M.T. Oxidative lime-pretreatment of high-lignin biomass—poplar wood and newspaper. Appl. Biochem. Biotechnol. 2001b;94(1):1–28.

Chenier P.J. Survey of industrial chemistry. second revised ed. New York: Wiley-VCH; 1992.

Gibson J., Gregory D.H. Carbonization of Coal. London: Mills and Boon; 1971.

Hammerschlag R. Ethanol's energy return on investment: a survey of the literature 1990–present. Environ. Sci. Tech. 2006;40:1744–1750.

Hirsch R.L., Bezdek R., Wendling R. Peak of world oil production and its mitigation. AICHE J. 2006;52(1):2–8.

Kroschwitz J.I., Howe-Grant M., Kirk R.E., Othmer D.F., eds. Encyclopedia of Chemical Technology. fourth ed. Hoboken, NJ: Wiley; 1991.

Lynd L.R. Overview and evaluation of fuel ethanol from cellulosic biomass: technology, economics, the environment and policy. Ann. Rev. Energy Environ. 1996;21:403–465.

McNeil D. Coal Carbonization Products. London: Pergamon Press; 1966.

Mokhatab S., Poe W.A., Speight J.G. Handbook of Natural Gas Transmission and Processing. Amsterdam: Elsevier; 2006.

Priyadarsan S., Annamalai K., Sweeten J.M., Muhktar S., Holtzapple M.T. Fixed bed gasification of feedlot manure and poultry litter biomass. Trans. ASAE. 2004;47(5):1689–1696.

Priyadarsan S., Annamalai K., Sweeten J.M., Holtzapple M.T., Muhktar S. Co-gasification of blended coal and feedlot and chicken litter biomass. Proc. Combust. Inst. 2005;30:2973–2980.

Ragauskas A.J., Williams C.K., Davison B.H., Britovsek G., Caimey J., Eckert C.A., Frederick Jr. W.J., Hallet J.P., Leak D.J., Liotta C.L., Mielenz J.R., Murphy R., Templer R., Tschalpliski T. The path forward for biofuels and biomaterials. Science. 2006;311:484–489.

Speight J.G. Chemical Process and Design Handbook. New York: McGraw-Hill; 2002.

Speight J.G. Natural Gas: A Basic Handbook. Houston, TX: GPC Books; 2007.

Speight J.G. An Introduction to Petroleum Technology, Economics, and Politics. Salem, MA: Scrivener; 2011a.

Speight J.G. The Refinery of the Future. Oxford: Elsevier; 2011b.

Speight J.G., ed. The Biofuels Handbook. London: The Royal Society of Chemistry; 2011c.

Speight J.G. The Chemistry and Technology of Coal. third ed. Boca Raton, FL: CRC Press; 2013.

Speight J.G. The Chemistry and Technology of Petroleum. fifth ed. Boca Raton, FL: CRC Press; 2014a.

Speight J.G. Gasification of Unconventional Feedstocks. Oxford: Elsevier; 2014b.

Speight J.G. Oil and Gas Corrosion Prevention. Oxford: Elsevier; 2014c.

Speight J.G. Handbook of Petroleum Refining. Boca Raton, FL: CRC Press; 2016.

Speight J.G., Islam M.R. Peak Energy—Myth or Reality. Salem, MA: Scrivener; 2016.

Sricharoenchaikul V., Frederick Jr. W.J., Agrawal P. Black liquor gasification characteristics. 2. Measurement of condensable organic matter (tar) at rapid heating conditions. Ind. Eng. Chem. Res. 2002;41:5650–5658.

Thanakoses P., Black A.S., Holtzapple M.T. Fermentation of corn stover to carboxylic acids. Biotechnol. Bioeng. 2003a;83(2):191–200.

Thanakoses P., Mostafa N.A.A., Holtzapple M.T. Conversion of sugarcane bagasse to carboxylic acids using a mixed culture of mesophilic microorganisms. Appl. Biochem. Biotechnol. 2003b;105:523–546.

Wilson Jr. P.J., Wells J.H. Coal, Coke, and Coal Chemicals. New York: McGraw-Hill; 1950.

Wittkoff H.A., Reuben B.G., Plotkin J.S. Industrial Organic Chemicals. third ed. Hoboken, NJ: Wiley; 2012.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset