Sie sind auf Seite 1von 106

CHAPTER TWO

LITERATURE REVIEW

2.1 2-Ethylhexanol

2-Ethylhexanol (2-ethyl-1-hexanol), C8H18O, ranks after the lighter alcohols

(methanol to butanol) as the most important synthetic alcohol. Approximately

2106 tonnes per annum are produced worldwide. 2-Ethylhexanol is mainly

used as the alcohol component for the manufacture of ester plasticizers for

soft poly(vinyl chloride) (PVC) and has been produced for this purpose since

the mid-1930s.

The raw material for the synthesis of 2- ethylhexanol, namely butyraldehyde

(butanal), is made almost exclusively from the petrochemical feedstock

propene via the oxo synthesis. The route from ethylene via acetaldehyde is

insignificant. This situation could change if acetaldehyde produced from

ethanol (via fermentation or homologation of methanol) and coal-based


synthesis gas (which is already used for the oxo synthesis) becomes available

at a reasonable price.

2.2 Properties of 2-Ethylhexanol

2.2.1 Physical Properties.

The straight-chain alcohols are colorless liquids at room temperature with

characteristic odors. The higher alcohols are solid, waxy substances. Because

of hydrogen bonding, the boiling points are considerably higher for alcohols,

especially lower-molecular-mass ones, than for the corresponding

hydrocarbons.With increasing molecular mass, however, the influence of the

hydroxyl group becomes less.

2-Ethylhexanol is a clear liquid with a characteristic odor and forms a

transparent mixture with other alcohols, ethers, and most organic liquids. Its

main physical properties are summarized in Tables 2.1 and 2.2


Table 2.1 Physical properties of 2-ethylhexanol (Falbe and Cornils, 1968)
Table 2.2 Temperature dependence of some physical properties of

2-ethylhexanol

2.1.2 Chemical Properties

2-Ethylhexanol reacts in the typical manner of -branched primary alcohols.

Reactions of alcohols can be characterized by cleavage of the OH bond or

the CObond either homolytically or ionically. The chemical properties

of greatest industrial importance are as follows.


Oxidation and Dehydrogenation.

Under normal conditions alcohols are stable. Oxidation with chemical agents,

such as chromic acid or permanganate, leads to a variety of products

depending on the nature of the alcohol.

Primary alcohols are oxidized first to aldehydes and then to carboxylic acids;

secondary alcohols are oxidized to ketones. Catalytic oxidation or

dehydrogenation of primary and secondary alcohols on copper, silver, iron,

molybdenum, etc., catalysts lead to the formation of aldehydes and ketones.

In the Oppenauer oxidation a secondary alcohol is dehydrogenated by an

excess of a ketone, e.g., acetone, in the presence of aluminum isopropoxide:

Reduction. With hydrogen iodide or zinc and hydrochloric acid, alcohols are

converted to hydrocarbons. Catalytic hydrogenolysis is especially successful

with benzyl alcohols.

Dehydration. Water can be split from alcohols by heating them in the

presence of strong acid or by passing them over aluminum oxide, silicic acid,

or synthetic zeolites. As a rule, not only those products that result from a -

elimination are formed, but also compounds with an isomerized double bond.
The isomerization can be suppressed by the addition of amines.

Tertiary alcohols can be dehydrated more easily than secondary or primary

alcohols. Under less severe conditions symmetric ethers are formed from

alcohols in an intermolecular reaction:

2ROH ROR + H20

The same compounds that accelerate the intramolecular dehydration are

suitable as catalysts.

Alcohols as Alkylation Reagents. Alcohols react with ammonia and amines

to form N-alkyl or N,N-dialkylamines. Aromatic hydrocarbons are alkylated

by alcohols in the presence of Friedel-Crafts catalysts.

Esterification. In the presence of acid catalysts, alcohols react with organic

or inorganic acids, acid chlorides, or anhydrides to form esters.

Addition Reactions. Alcohols add to aldehydes and ketones to form acetals.

Alkyl polyglycol ethers are obtained with alkylene oxides. The addition to

acetylene gives vinyl ether; addition to olefins yields mixed ethers.

2.3 Industrial Production of 2-Ethylhexanol

2-Ethylhexanol is produced in four steps:


1) Aldolization of butyraldehyde and subsequent dehydration

2) Separation of the aldolization solution

3) Hydrogenation of unsaturated 2-ethyl-2-hexenal as an intermediate product

4) Fractionation of 2-ethylhexanol

The butyraldehyde required for the aldolization stage is made by

hydroformylation of propene. In addition to processes using cobalt as a

catalyst (Cornils and Falbe, 1980), processes that employ rhodium catalysts

(Matthey, 1977) have gained increasing significance since the 1980s.

The demands placed on the purity of 2-ethylhexanol are so high that the purity

of the starting product, butyraldehyde, is also extremely important.

Isobutyraldehyde is formed to a greater (cobalt catalysts) or lesser (rhodium

cata-lysts) extent during hydroformylation and must be separated to prevent

mixed aldolization.

Therefore, special processes in which aldolization and, in some cases,

hydrogenation occur in parallel with hydroformylation are becoming less


important (Cornils, and Falbe, 1980; Shell, 1972).

Older processes based on acetylene that involve the intermediate stages

acetaldehyde/ crotonaldehyde would permit access not only to coal as a raw

material, but also to petrochemistry via ethylene Weber et at al., (1976),

Kyowa, (1967) as well as to regenerable raw materials via ethanol. However,

these processes are not currently important.

The major oxo producers ( BASF, 1973; Ruhrchemie, 1985; UCC, 1975) all

offer processes for the manufacture of 2-ethylhexanolthat differ only in certain

details. Figure 2.1 shows the flow diagram for a 2-ethylhexanol plant.

Figure 2.1 2-Ethylhexanol plant

a) Aldolization; b) Phase separator; c) Hydrogenation; d) Distillation

In the aldolization reactor, butyraldehyde reacts very quickly to give 2-ethyl-

2-hexenal, aqueous sodium hydroxide being employed as the standard

industrial catalyst. Local overheating in the reaction mixture must be avoided,

since this may cause secondary reactions and thus decrease yields. Thorough

mixing of the two-phase system is necessary. The primary aldol addition must
take place rapidly; the immediately ensuing dehydration of the hydroxyalde

hyde must be conducted as quickly and quantitatively as possible because the

aldol is unstable and can impair the product quality and yield.

The aldolization reactor may be a mixing pump (Ruhrchemie, 1976) a packed

column (Dumbgen, 1969), or a stirring vessel(Kuhlmann, 1968). The various

processes operate at a temperature of between 80 and 150 C and pressures

below 0.5MPa. The ratio of aldehyde to aqueous sodium hydroxide solution

is 1 : 10 1 : 20.

Under these conditions, conversion rates>99% are obtained. The heat of the

aldolization reaction is used for steam generation.

After the aldolization stage in (a) (see Fig. 2.1), the mixture is separated in a

phase separator (b) into an upper organic phase and a lower aqueous phase

containing the aldolization solution.

Part of the aldolization solution can be recycled, but the rest must be removed

from the system via a side stream because the aldolization solution is diluted

by the water that is produced in the reaction; the aldolization byproducts

(Cannizzaro and Tischchenko reactions) must also be removed. Sodium

hydroxide is added to maintain its concentration at 2 4 %.

The side stream has such a high COD value (e.g., due to its sodium butyrate

content) that it should be subjected to preliminary purification before being


conveyed to a biological treatment plant. Suitable purification methods are

oxidation, acid treatment, and extraction which allows partial recycling of

valuable products.

The organic product from the phase separator (b) can either be hydrogenated

in a single stage on a fixed nickel catalyst (Matthey, 1977) or in several stages

either in a combination of gas liquid phases or liquid liquid phases (sump

phase or trickle-bed reactor). The heat of reaction for the hydrogenation of the

C=C double bond and the aldehyde group is relatively high, 178 kJ/mol.

Temperature control problems may therefore arise as local overheating

decreases yields and must be avoided.

With single-step hydrogenation, remixing with the hydrogenation product has

been proposed to dissipate heat (150 200 C)(Dumbgen, 1969); in contrast

to other processes, medium pressure is initially necessary to ensure adequate

conversion.

Modern plants normally utilize two stages to remove residual amounts of

carbonyl compounds and to ensure that high-grade 2-ethylhexanol is obtained

(Cornils, and Mullen, 1980). Nickel (Dumbgen, 1969), copper or mixed

systems. The heat of the hydrogenation reaction is also used for steam

generation. In the hydrogenation stage, a conversion of 100 % and a selectivity


of >99 % is attained.

Fractional distillation of the hydrogenation product from (c) in (d) normally

takes place in three stages: In the first stage, the light ends are separated at the

head and can be employed for the manufacture of 1-butanol. In the second

stage, pure 2-ethylhexanol is collected at the head. In the third stage, the

recyclable intermediate fractions are separated from the heavy oil. The

byproducts may be used for heating purposes.

Trade Names. 2-Ethylhexanol is usually marketed by the manufacturers (e.g.,

Ruhrchemie, BASF, Huls, Union Carbide, Mitsubishi Chemical) under its

chemical designation 2-ethyl-1-hexanol or as octyl alcohol.

Typical specifications have been compiled in Table 2.3.

Table 2.3. Quality specifications for 2-ethylhexanol

2.4 Storage and Transportation. Safety precautions for the shipment and

storage of 2- ethylhexanol are determined by its combustibility, flash point,


and ignition temperature (seeTable 4). Dry 2-ethylhexanol does not corrode

standard metals and, like other alcohols, can be stored in standard steel

containers (e.g., ST 37).

For particularly high-quality demands, use of stainless steel or aluminum

containers is advisable.

To prevent the entry of atmospheric oxygen or moisture, an inert gas cover

should be provided. 2-Ethylhexanol can be transported in iron drums, tank

trucks, or railway cars made of standard steel, aluminum, or stainless steel.

Centrifugal pumps conforming to chemical standards with simple rotating

mechanical seals and polytetrafluoroethylene (PTFE) flange seals have

proved to be suitable for use with 2-ethylhexanol.

Table 2.4. Safety information for 2-ethylhexanol

2.5 Uses of 2-ethylhexanol

In the United States, 2-ethylhexanol is mainly used in the form of its secondary
products for the following applications (CMR, 1984):

60 % share of the poly(vinyl chloride) PVC) plasticizers market. In the United

States, bis(2-ethylhexyl) phthalate [117-81-7], di-2-ethylhexyl phthalate

(DEHP), also called dioctyl phthalate (DOP), dominates (70 %); this is

followed by bis(2-ethylhexyl) adipate [103-23-1], di-2-ethylhexyl adipate

(DEHA), also called dioctyl adipate (DOA), (16%), and tris(2-ethylhexyl)

mellitate, trioctyl trimellitate (12 %). In Europe, DEHPs share of the market

is even higher. The outstanding importance of DEHP is due to its favorable

property combination as a plasticizer (good gelling properties, low volatility,

high resistance to heat and water, and excellent electrical properties).

Comprehensive investigations have been conducted on the toxicology and

ecology of the widely used chemical DEHP during the last few years. The

results have shown that its health and ecotoxicological risks are extremely

low(ECETOC, 1984).

The second large field of application is the production of 2-ethylhexyl acrylate

[103-11-7], which is used to manufacture coating materials (especially

emulsion paints), adhesives, printing inks, impregnating agents, and reactive


diluent/cross-linking agents. 2-Ethylhexyl nitrate [27247-96-7] serves as a

cetane number improver (Ethyl Corp, 1984). 2-Ethylhexyl phosphates are

used as lubricating oil additives (Elco Corp, 1985). Antifoaming agents,

dispersants, and flotation agents made from 2-ethylhexanol deserve special

mention as surfactants. Other applications of 2-ethylhexanol include its use as

a solvent, above all for polymerization catalysts and in extracting agents.

2.6 Toxicology

Oral LD50 values of 2-ethylhexanol are ca. 2.0 g/kg and 3.7 g/kg in rats. Gross

necropsy revealed evidence of gastrointestinal irritation. In the rabbit eye, the

undiluted material produces an erythema, swelling of the conjunctiva,

lacrimation, and mucous secretion.

After skin contact in rabbits, 2-ethylhexanol is mildly irritating if the patch is

not closed; otherwise moderate skin irritation occurs.

The lowest published LD50 after dermal application is 2 g/kg in rabbits.

Mice, rats, and guinea pigs survived inhalation of a nearly saturated

atmosphere (227 mL/m3) over a period of 6h. Typical signs of intoxication

were central nervous depression, dyspnea, and irritation of mucous

membranes. Repeated inhalation of a maximum dose of 110 mg/m3 by rats

caused dystrophic changes in parenchymatous organs.

The cumulative oral toxicity seems to be low. A total of 12.5 g/kg in the diet
fed to rats over a period of 90 d affected the kidneys and the liver. An amount

of 2 mL/kg fed to rabbits for 10 d caused signs of local irritation and systemic

toxicity. Peroxisomes were induced with liver enlargement when rats were fed

with dietary levels of 2% 2-ethylhexanol for a period of 3 weeks.

In rabbits and rats, 2-ethylhexanol is conjugated to glucuronic acid and

eliminated via the urine, its main metabolite being 2-

ethylhexanonylglucuronide. A small amount of the radioactively labeled

material is oxidized and expired as CO2. Only 2 % is excreted unchanged.

Neither the ACGIH, OSHA, nor DFG have set standards for exposure levels.

2.7 HYDROFORMYLATION

Hydroformylation, often referred to as oxo synthesis, was discovered by Otto

Roelen in Germany before World War II in one of the many programs aimed

at the use of coal-derived synthesis gas (CO/H2 mixtures). Roelen discovered

that alkenes react with syngas, provided an appropriate catalyst is present.

Formally, a formyl group (CHO) and a hydrogen atom are added to the double

bond. Therefore, the reaction has been called hydroformylation, analogous

to hydrogenation being the addition of hydrogen. The most important

products are in the range C4C19. With a share of about 75%, the most

significant is the hydroformylation of propene (Bahrmann and Bach, 2000)

yielding two isomers:


The products of hydroformylation usually are intermediates for the production

of several types of alcohols, which are commonly known as oxo-alcohols. In

most cases the normal aldehyde is the preferred product because it enjoys a

much larger market than the branched aldehyde. For instance, normal butanol,

produced by the direct hydrogenation of normal butyraldehyde, is used in the

production of a wide variety of chemical intermediates, while iso-butanol is

predominantly used as a solvent. Furthermore, only normal butyraldehyde can

be used to produce 2-ethylhexanol, which is the most widely used plasticizer

alcohol (Tudor and Ashley, 2007). The oxo-alcohols in the C11C17 range are

utilized in detergents.

2.7.1 Thermodynamics of hydroformylation

Hydroformylation requires low temperature and elevated pressure, as can be

seen from Figure 2.2a, which shows the equilibrium conversion of propene to

the corresponding aldehydes, at a typical molar inlet composition


H2/CO/propene = 1:1:3. As mentioned earlier, in general, normal aldehydes

are the preferred products.

However, it is clear from the equilibrium data in Figure 2.2b that the iso-

aldehyde is the thermodynamically favored product. During hydroformylation

also other reactions can occur, in particular hydrogenation to propane.

Thermodynamically, the latter reaction is the preferential reaction, so a

catalyst with high selectivity is required.

Figure 2.2 Hydroformylation of propene: (a) propene equilibrium conversion; (b) normal/iso

distribution of butyraldehyde (p = 30 bar; propene/syngas = 3 mol/mol).

2.7.2 Catalyst Development

The most important catalysts for hydroformylation are based on rhodium and

cobalt, which are introduced as carbonyls. Rhodium and cobalt are not the

only metals showing catalytic activity but they are the most active ones. The
relative metal activity for hydroformylation is shown in Table 2.5. It is not

surprising that in practice rhodium and cobalt are used.

A wide variety of ligands, in particular phosphines (Figure 2.3, have been

shown to influence activity, stability, and selectivity of the catalyst complexes.

The first catalyst used for the hydroformylation reaction, in the 1940s, was

cobalt hydridocarbonyl, HCo(CO)4, which requires operation at very high

pressure (200450 bar) to ensure catalyst stability (Tudor and Ashley, 2007).

In the 1960s, a more stable tributylphosphine-substituted cobalt carbonyl

catalyst, HCo(CO)3P(n-C4H9)3, was developed, which also shows a much

higher normal-toiso ratio in the product distribution. However, the activity of

the catalyst is lower and more by-products are formed (Billig and Bryant,

2000).

Table 2.5 Relative activities of metals for hydroformylation.

Rh Co Ru Mn Fe Cr,Mo,W,Ni

103104 1 102 104 106 0


Figure 2.3 Examples of phosphine ligands.

The continuing search for catalyst systems that could effect the

hydroformylation reaction under milder conditions and produce higher yields

of the desired aldehyde resulted, in the 1970s, in processes utilizing rhodium.

The active catalyst used in most processes is a rhodium complex modified

with triphenylphosphine (TPP) ligands, HRh(CO)(PPh3)3. This catalyst

system has a high activity, stability, and selectivity (Arnoldy, P. (2002). Since the

1980s, a rhodium catalyst with a water-soluble ligand, triphenylphosphine-m-

trisulfonic acid trisodium salt (TPPTS) has also been used. The normal-to-iso

ratio of the product is very high, but the activity of this catalyst is significantly

lower (Billig and Bryant, 2000).

Kinetic studies have been performed for the hydroformylation reaction. For a

simple cobalt system, (HCo(CO)4), the following rate expression has been

found (Chaudhari, 2001):

For a rhodium system containing TPP ligands the following rate equation has

been reported (Van Koten, and Van Leeuwen, 1999):


During hydroformylation, side reactions occur, for example, double bond

isomerization (e.g., 1-hexene to 2-hexene) and hydrogenation of the alkenes

to the corresponding alkanes. The extent of these side reactions depends on

the metal and the ligand added.

The selectivity to normal aldehydes of rhodium catalysts is much better than

that of cobalt catalysts. In addition, rhodium catalysts show little

hydrogenation or double bond isomerization activity. Therefore, processes for

hydroformylation of propene, the predominant feedstock, are currently almost

exclusively based on this type of catalyst.

Fig. 2.4 illustrates the catalytic cycle for rhodium-based hydroformylation.

Catalytic Cycle for Rhodium-Based Hydroformylation shows the catalytic

cycle for hydroformylation of propene by HRh(CO)L2 (Billig, and Bryant,

2000).
Fig 2.4 Catalytic cycle in hydroformylation of propene; L =

triphenylphosphine ligand.

Unlike rhodium catalysts, cobalt catalysts strongly enhance double bond

isomerization. This is why cobalt based processes are quite flexible in that

they can be used for the hydroformylation of a wide variety of alkenes,

including internal alkenes, which would yield a large proportion of the

undesired nonlinear products with a rhodium catalyst. fig 2.5 illustrates how

double bond isomerization increases the selectivity to linear aldehydes.


Fig. 2.5 Example of the hydroformylation of internal alkenes with a

cobalt catalyst.

2.7.3 Processes for the Hydroformylation of Propene

2.7.3.1 Low-Pressure Oxo Process

Hydroformylation of lower alkenes, in particular propene, nowadays is

usually based on rhodium catalysts. By adding the right ligands the selectivity

towards linear products is high. The reaction takes place in a continuous

stirred-tank reactor at much lower pressure than the processes with

unmodified cobalt catalyst. The alkene and synthesis gas have to be purified

thoroughly because the catalyst is extremely sensitive to poisons. Figure 2.6

shows a process scheme for the rhodium-catalyzed hydroformylation of

propene. This process is often called the low-pressure oxo process;

information about its development is given elsewhere (Tudor, R. and Ashley,

2007).

The gaseous reactants are fed to the reactor through a sparger. The reactor
contains the catalyst dissolved in product butyraldehyde and by-products (e.g.,

trimers and tetramers). Rhodium stays in the reactor, apart from a small

purification cycle.

Figure 2.6 Rhodium-catalyzed hydroformylation of propene (low-


pressure oxo process).

The gaseous reactor effluent passes through the demister, in which fine

droplets of catalyst that are contained in the product gases are removed and

fed back to the reactor. The reaction products and unconverted propene

(conversion per pass is about 30%) are condensed and fed to the gas/liquid

separator. Propene, apart from a small purge stream, is recycled to the reactor.

The liquid reaction products are freed from residual propene in a stabilizer

column and further purified by distillation.


A portion of the catalyst solution passes through a purification cycle for

reactivation of inactive rhodium complexes, and for removal of heavy ends

and ligand decomposition products.

The demisting device is an essential part of the plant. It ensures that all

rhodium remains in the reactor. In view of the high cost of rhodium, recovery

should be as high as possible. In practice, the rhodium losses in the process

are very small.

2.7.3.2 Process with a Biphasic Catalyst System

A novel industrial process is based on a water-soluble rhodium catalyst

(Ruhrchemie/Rhone-Poulenc, now Celanese) (Kohlpaintner, et bal., 2001). It

contains polar ligands (sulfonated triphenylphosphines) which are highly

soluble in water.

The catalyst is insoluble in the organic phase formed, so a biphasic reaction

medium results. Thus, separation is greatly simplified and, as a consequence,

the rhodium losses are minimal. In fact, the catalyst system could be

considered to be heterogeneous. Figure 2.7 shows a simplified flow scheme

of the process.

In this process, the reaction also takes place in a continuous stirred-tank

reactor containing the catalyst solution. Before entering the reactor, the syngas
is first passed through a stripping column to recover the unreacted propene.

The liquid reactor effluent is fed to a phase separator where dissolved gases

are removed and the butyraldehydes are separated from the aqueous catalyst

solution. The catalyst solution is returned to the reactor via a heat exchanger

in which steam is generated.

Figure 2.7 Ruhrchemie/Rho ne-Poulenc hydroformylation of propene

using a water-soluble rhodium catalyst.

2.7.4 Processes for the Hydroformylation of Higher Alkenes

2.7.4.1 Conventional Cobalt-Based Process

The first industrial oxo process was based on an unmodified cobalt catalyst:

HCo(CO)4. This catalyst system requires high pressures and temperatures to

ensure the stability of the catalyst. Hydroformylation of higher alkenes to

produce plasticizer and detergent-range alcohols still uses this type of catalyst.
To minimize catalyst consumption and to avoid problems in downstream

processing (fouling of equipment,etc.), it is very important that cobalt is

recovered. A process that offers an elegant solution to the problem of catalyst

separation is the Kuhlmann (now Exxon Mobil) process. This

hydroformylation process is based on the following cobalt cycle (Billig and

Bryant, 2000):

2HCo(CO)4 + Na2CO3 2NaCo(CO)4 + H2O + CO2

The sodium salt is soluble in water. This enables recovery of the cobalt catalyst

by scrubbing with water. The aqueous solution containing cobalt as the

sodium salt is transformed to the active catalyst by reaction with sulfuric acid:

2NaCo(CO)4 + H2SO4 2HCo(CO)4 + Na2SO4

This complex is soluble in alkenes and, as a consequence, it can be returned

to the reactor dissolved in the reactant alkene flow.

Figure 2.8 shows a flow scheme of the Kuhlmann hydroformylation process

for the production of aldehydes. The alkene, with recycled and make-up

catalyst, is fed to the hydroformylation reactor (usually a stirred tank or a loop

reactor) together with the synthesis gas.


Figure 2.8 Kuhlmann hydroformylation process for the production of

aldehydes.

The crude product, after passing through the flash vessel, in which the gas is

separated, is treated in countercurrent flow with an aqueous sodium carbonate

(Na2CO3) solution. The cobalt complex forms a water-soluble sodium salt.

After scrubbing with water in the wash column, a virtually cobalt-free organic

phase (crude aldehydes) and an aqueous phase containing the sodium salt

remain. The crude aldehydes are sent to a distillation section (not shown).

Here the excess alkenes are recovered and recycled to the reactor. The sodium

salt is transformed to the original cobalt complex, HCo(CO)4, by adding

sulfuric acid in the regenerator. This complex is extracted by the alkene

feedstock and recycled to the reactor.


2.7.4.2 Low-Pressure Cobalt-Based Process for the Direct Production of

Alcohols

Shell has commercialized a process based on a trialkylphosphine-cobalt

catalyst that is much more stable than the unmodified catalyst, allowing (and

requiring) operation at lower pressure (Bahrmann and Back, 1991). This

process produces the alcohol rather than the aldehyde, which is an added

advantage since alcohols are usually the desired end products of

hydroformylation.Adisadvantage is that the high hydrogenation activity also

causes the formation of alkanes (up to 15% versus 2% for the non-modified

catalyst). Therefore, it may be advantageous to use two reaction stages, with

a lower hydrogen partial pressure in the first reactor to promote

hydroformylation of the alkene rather than hydrogenation. Figure 9.11 shows

a simplified flow scheme of this version of the process.

The alkene feed is added to the first reactor together with make-up and recycle

catalyst complex. Here, conversion takes place with a hydrogen-poor

synthesis gas to limit alkane formation. In the second reactor, hydrogen-rich

gas is added to ensure the hydrogenation of the aldehyde to the alcohol. The

separation takes place by depressurization in a flash vessel and subsequent

distillation. In the first column, unconverted alkenes are removed and recycled

to the reactor. A purge stream is added in order to prevent the build-up of


alkanes. The second column separates the catalyst complex and heavy by-

products from the crude alcohols.

Figure 9.11 Shell hydroformylation process using a modified cobalt


complex.

By-products are formed by dimerization of aldehydes. Build-up of these aldols

is prevented by purging part of the recycle stream.

Although this process looks much simpler than the Kuhlmann

hydroformylation process, it has two distinct disadvantages:

thorough purification of the synthesis gas is very important because the

ligand is very sensitive to oxidation;

the activity of the modified cobalt catalyst is much lower than that of the

unmodified catalyst and, therefore, for the same productivity the reactors

need to be five to six times larger than in the conventional process.


Advantages, apart from the ones already mentioned, are:

higher linearity of the alcohol product;

the possibility to separate the catalyst from the product by conventional

distillation.

2.7.5 Comparison of Hydroformylation Processes

Table 2.6 summarizes characteristics of the major hydroformylation

processes. The rhodium catalyst is attractive for the conversion of propene to

the linear aldehyde. The selectivity may be as high as 95%.

The branched product also has a market, although a much smaller one than

the linear product, and its production is, therefore, desired in these small

quantities. Hydroformylation of higher alkenes with internal double bonds

requires isomerization activity and, as a consequence, in that case cobalt

catalysts are preferred. In all processes, except one, aldehydes are the primary

products. These can be hydrogenated to alcohols in a separate reactor. The

modified cobalt catalyst is used for the direct production of alcohols rather

than aldehydes because this catalyst promotes hydrogenation.

Table 2.8 Comparison of hydroformylation processes (Chauvel and


Lefebvre, 1989; Falbe, 1980; Bahrmann and Back, 1991).
2.8 SYNGAS

Synthesis gas (syngas) is a mixture of hydrogen, carbon monoxide and carbon

dioxide. It may also contain nitrogen as applied for the ammonia synthesis.

Syngas is a key intermediate in the chemical industry. It is used in a number

of highly selective syntheses of a wide range of chemicals and fuels, and as a

source of pure hydrogen and carbon monoxide. Syngas is playing an

increasing role in energy conversion (Rostrup-Nielsen,2004).

Synthesis gas can be produced from almost any carbon source ranging from

natural gas and oil products to coal and biomass by oxidation with steam and

oxygen. Hence it represents a key for creating flexibility for the chemical

industry and for the manufacture of synthetic fuels (synfuels).


Figure 2.8 Conversion via syngas.

The conversion via syngas results in products plus heat (Figure 2.8). In most

plants, the heat is utilised for running the plant. As an alternative, the heat may

be exported, but that is not always necessary.

The present use of syngas is primarily for the manufacture of ammonia (in

2006, 124 million tonnes per year) and of methanol (in 2005, 33 million tonnes

per year), followed by the use of pure hydrogen for hydrotreating in refineries

as shown in Table 2.6.

The main commodity products based on natural gas are shown in Table 1.1

(Rostrup-Nielsen, 2005). It is evident that the chemical conversion of natural

gas (approximately 7 109 GJ/y) is marginal to the total natural gas production

(3.07 1012 Nm3/y [78] or 1.17 1011 GJ/y assuming a lower heating value

(LHV) equal to 38 MJ/Nm3). Recent trends in the use of syngas are dominated

by the conversion of inexpensive remote natural gas into liquid fuels (gas to

liquids or GTL) and by a possible role in a future hydrogen economy

mainly associated with the use of fuel cells. These trends imply, on the one

hand, the scale-up to large-scale GTL plants (more than 500,000 Nm3
syngas/h) and, on the other hand, the scaledown to small, compact syngas

units for fuel cells (5100 Nm3 syngas or H2/h). These forecasts create new

challenges for the technology and for the catalysis.

Table 2.6 Main chemical products based on natural gas.

The data in Table 2.6 show that the practical efficiencies for natural gas

conversion into products are approximately 80% of the ideal values expressed

as:

ideal = LHV methane/mol


LHV product /mol (1.1)

For endothermic reactions (ethylene, hydrogen), the LHV of the fuel

providing the reaction heat should be added to the nominator.

The world energy production is dominated by fossil fuels as the main energy

source. It amounted to 88% in 2007 with oil responsible for 37% become the
worlds second largest consumer of oil, after the USA. The proved reserves of

oil are concentrated in the Middle East (61%) and those of natural gas are also

in the Middle East (41%), followed by Russia (23%) (BP, 2009). Coal is more

evenly distributed between the continents.

Apart from the large reserves (Middle East, Russia), natural gas is present as

associated gas in oil fields. However, as many fields are far from the

marketplace and often off-shore, the gas there is called remote gas or stranded

gas (Chew, 2003). Part of the associated gas is reinjected to enhance the oil

recovery, but unfortunately still a significant fraction is flared for

convenience. The flared gas amounts to close to 5% of the total natural gas

production (corresponding to about 1% of total world CO2 production from

fossil fuels).

So far, the proven reserves for oil have followed the increase in production as

expressed by the reserves/production ratio (R/P ratio) staying at about 40 for

oil over the last 20 years; however, at a steadily increasing cost of exploration

and production. A big fraction of the reserves is present as oil sand (tar sand)

and other non-conventional sources under active development (BP, 2009).

This means that at the present world production, the oil reserves known today

will be used up within about 40 years. This figure should be considered with
care. It does not include reserves still to be found and it does not include the

changes in consumption (for instance the growth in Asia). Furthermore, the

R/P ratio for oil varies from region to region, being above 80 in the Middle

East and below 20 in North America.

The R/P ratio (2007) for natural gas is about 60 and 122 for coal(BP, 2009).

The total R/P for fossil fuels (based on oil equivalent) is less than 100 years.

These figures emphasise the need for flexibility in the energy network and the

need for alternative fuels. Oil is the most versatile of the fossil fuels with high

energy density and it is easily transported.

The power industry is very flexible to feedstocks and it is feasible to transport

coal over long distances to big centralised power plants close to deep water

harbours. Natural gas is transported to the marketplace in pipelines over still

longer distances or as liquified natural gas (LNG).

The automotive sector represents a special challenge as the energy conversion

is strongly decentralised. So far oil-derived products have been the solution,

but in view of the limited reserves of oil, a number of alternative fuels are

being considered, such as liquefied petroleum gas (LPG), natural gas,

methanol, dimethylether (DME), ethanol, biodiesel, synfuels and hydrogen.

Biofuels represent a sustainable response to liquid fuels. It may be based on


ethanol and biodiesel derived from conventional agricultural products or from

synfuels via gasification of biomass. The alternative fuels may be blended

with conventional fuels or used directly in internal combustion engines (ICE)

or fuel cells. In Western Europe alternative fuels may amount to 20% of energy

sources by 2020.

Globalisation has caused companies to concentrate on core business and

critical mass. It has resulted in a restructure of the chemical industry into two

types of focused companies (Felcht, 2003): the molecule suppliers

(commodities and fine chemicals) and the problem solvers (functional

chemicals like additives and pharmaceuticals). Each type has its own

characteristics as reflected by the role of the catalyst (Rostrup-Nielsen, 2004).

The most important parameter for large-volume chemicals is production costs

(variable and fixed costs). The variable costs are related to the feed costs, the

use of energy, process selectivity and environmental costs.

Four trends have characterised plants for commodity chemicals:

Location of cheap raw materials;

Economy of scale;

More integrated plants; and

CO2 footprint (tonnes CO2 per tonne product).


Plants are moved to locations where raw materials are cheap. As illustrated in

Figure 2.9, the ammonia production is hardly feasible at natural gas prices

typical for Europe and USA (34 USD/GJ with high seasonal variations)

(Rostrup-Nielsen, 2005). As a result, new plants for commodity chemicals are

built at locations (Middle East, Trinidad, Nigeria, West Australia) with low

natural gas prices (0.51 USD/GJ). It means that the use of natural gas as

feedstock may not be feasible where there is a big market for natural gas as

fuel.

Figure 2.9 Ammonia production costs (Rostrup-Nielsen, 2005).


Reproduced with the permission of Springer.

For any process scheme, it is essential at an early stage to establish the overall

mass balance and to estimate the P as simply being the difference between

the price of products and the price of feedstocks (Rostrup-Nielsen, 2004).


Hence, there has been a trend to develop processes using cheaper raw

materials. The gain in P could, however, be lost by lower selectivity or higher

investments. Selectivity is crucial to achieving a high P. Low selectivity and

conversion per pass result in low concentrations in process streams and hence

more expensive separation systems.

Figure 1.3 Simplified mass balance.

It may be argued that energy efficiency is of less importance when natural gas

is cheap, but high energy efficiency means small feed pretreat units and

reduced requirements for utilities and hence less investments.

Moreover, high efficiency means less CO2 production. As illustrated in Figure

1.3, the P calculation should consider also the energy consumption and the

by-products which may easily have a negative value. This may be expressed
by the so-called E-factor (Sheldon, 1997) expressing the amount of by-product

produced per kg of product.

The emissions may have great negative value. This may be reflected by the

costs of carbon capture and storage (CCS).

The CO2 emission expressed as a C-factor (Rostrup-Nielsen, 2005) (tonnes

CO2 per tonne product, refer to Table 1.1) may become an important process

parameter in the future. CO2 emissions are often directly related to the energy

consumption of the process.

As an example, a reduction of the energy required to produce ammonia from

natural gas of 1 GJ/t means a reduction of CO2 emissions of around 8.5

million tonnes CO2/y, worldwide. In many ammonia plants about 80% of the

CO2 is reacted with ammonia to urea from which it is, of course, liberated to

the atmosphere when the urea is used as fertiliser.

On the other hand, CO2-consuming processes will hardly change the picture

from non-carbon containing fuel.

As an example, consider the methanol synthesis:

CO2 3H2 CH3OH H2O


Even if hydrogen was made available from non-carbon containing fuel, the

present world production of methanol via this reaction would only amount to

40 million t CO2/y. This corresponds to the CO2 emission from a 4000 MW

coal-based power plant and should be compared with the total CO2 emissions

of approximately 27.5 109 t CO2/y (7.5 109t carbon/y). It means that CO2 as

a reactant will have little impact on the CO2 problem. Again, the products will

eventually end as CO2. A similar argument is valid for CO2 reforming of

natural gas.

2.8.1 Direct or indirect conversion

An important challenge in C1 chemistry is to circumvent the syngas step by

a direct conversion of methane into useful products. Still, yields are far from

being economical (Holmen, 2009; kuo et al., 1989). The methane molecule is

very stable, with a C-H bond energy of 439 kJ/mol; hence methane is resistant

to many reactants. Electrophillic attack requires superacidic conditions, and

radical abstraction of a hydrogen atom by a reactant Q requires that the Q-H

bond exceeds 439 kJ/mol (Crabtree, 1995):

CH4 + Q_ CH3+ QH

This is feasible when Q is an oxidising agent. However, the product often has

much weaker C-H bonds than methane, which implies that it is difficult to

eliminate further reactions leading to complete oxidation.


CH4 O 2
CHxO O CO2 + 2H2O
2

The direct conversion of CH4 into methanol may have a high selectivity, but

at a low conversion per pass. For example, Zhang et al. (2003) reported a

selectivity of 60% at a conversion of 1213%. This corresponds to a yield of

about 7.5%. This low yield per pass results in a large recycle ratio and a

difficult separation associated with a low partial pressure of the product. This

is illustrated by simple calculations in Figure 2.10 (Rostrup-Nielsen, 2005).

Figure 2.10 Recycle ratio and conversion. A selectivity of 95% at a conversion per pass of 5%
means a large recycle ratio of approximately 10, and hence a difficult separation due to low partial
pressure of the product. Reproduced with the permission of Springer [410].

A simple kinetic analysis of consecutive first-order reactions (Rostrup-

Nielsen, 2005) may illustrate the problem. The data in Figure 2.11 show that
the higher the ratio k2/k1, the lower the yield of the intermediate B.

Figure 2.11 Consecutive reactions. Reactivities and maximum yields. Reproduced with the
permission of Springer .

The direct oxidation of methane to methanol or formaldehyde has been a

dream reaction for a long time (Zhang, 2003). Attempts include gasphase

reaction, catalytic reactions, and use of other oxidants than air.

Selectivities may be high, but at a lower conversion per pass resulting in yields

being inferior for industrial use.

High selectivity and conversion may not be sufficient. A process using

superacid activation (Catalytica) for converting CH4 via methyl bisulphate

into methanol has the potential of achieving a high selectivity of 95% at a


conversion of 90% (Periana, 1998). However, the process would require a

large sulphuric acid plant (1 mol SO3/mol methanol) and a unit for

concentrating a large recycle of diluted acid.

Other attempts have aimed at creating a carbon-carbon bond from methane,

although most natural gas sources contain a fraction of ethane and other lower

alkanes.

Most work in direct conversion has focused on the oxidative conversion of

methane into ethylene (Fredenslund et al., 1977). It has proven to be more

promising than high-temperature pyrolysis of methane into primarily

acetylene. However, the process suffers from ethane being a significant part

of the products (low P) and that above 20% of the converted methane is

oxidised to carbon oxides. Under industrial conditions C2+ yields are less than

20% at a conversion of 2435% per pass. As a result, the process scheme ends

up being rather complex, meaning that the oxidative coupling is not

economically feasible with the present low selectivities to C2 hydrocarbons.

Although the reaction scheme is elegant, the principles behind Figure 1.4 may

explain why yields in oxidated coupling have never passed an apparent ceiling

( Lee, 1988).

Catalytic partial oxidation at high temperature and ultra-short residence time


over noble metals gauze has shown formation of olefins and oxygenates

(Goetsch, 1996). The feasibility of this route is still to be analysed. The

indirect route via methanol appears to be a more promising route for olefins.

Direct conversion of methane to higher hydrocarbons without the assistance

of oxygen is not favoured by thermodynamics. This constraint can be

circumvented in a two-step process via carbides, but so far yields have been

insignificant (Kjr, 1972).

Other studies have explored the direct conversion of methane into benzene.

Selectivities of 70% were obtained close to equilibrium conversion at 600oC

(12%).

From a thermodynamic point of view the manufacture of synthetic

transportation fuels should aim at a minimum change of the hydrogen content

of the feedstock to that of the product (typically around H/C=2). It means that,

in principle, it is more efficient to convert natural gas to paraffinic diesel than

to aromatic rich gasoline. For coal the

indirect conversion via syngas appears less efficient than the direct

hydrogenation routes. However, these theoretical considerations should be

supplemented with an analysis of the process steps and selectivities involved.

The main advantage of the indirect routes via syngas is the very high carbon
efficiency. As an example, a modern methanol synthesis loop based on natural

gas may operate with more than 50% conversion per pass having a selectivity

of 99.9% and a carbon efficiency above 95%. The synthesis gas routes are

highly efficient, but they are capital intensive because they involve exchange

of energy in the reformers and heat recovery units, as illustrated in Figure 2.12

(Rostrup-Nielsen, 2002).

Figure 2.12 Indirect conversion of natural gas (numbers indicate the


relative investments) (Rostrup-Nielsen, 2002). Reproduced with the
permission of Elsevier.

Syngas manufacture may be responsible for approximately 60% of the

investments of large-scale gas conversion plants based on natural gas.

Therefore, there is great interest in optimising process schemes based on

steam reforming and autothermal reforming as well as in exploring new routes


for the syngas manufacture.

2.9 Propene 15967

Propene, CH3CH=CH2, Mr 42.081, [115-07-1] was the first petrochemical

raw material to be employed on an industrial scale and was used more than 60

years ago in the production of isopropanol. For a long time, propene was to

some extent overshadowed by its olefin homologue ethylene. However, part

of propenes turbulent development and expansion since 1965 is also due to

the success of ethylene: Being a byproduct of ethylene production, many

important areas of application were opened up to propene by the chemical

industry. Secondary products of propene, such as polypropylene, acrylonitrile,

propylene oxide, and in Europe cumene, have now clearly overtaken the

classic secondary product isopropanol in importance.

2.9.1 Physical Properties


Table 2.7 physical properties of propene:

The temperature dependence of typical properties is given in Table 2.8.

2.9.2. Chemical Properties

The chemical properties of propene are, like those of ethylene, characterized


by the reactivity of its double bond. Propene undergoes a number of

industrially important polymerization, addition, and oxidation reactions.

Table 2.8. Temperature dependence of some physical properties of

propene, according to (Yaws,1975).

2.9.3 Production of Propene

Although propene is one of the most important feedstocks for the organic

chemicals industry, it is produced almost entirely as a byproduct because it is

obtained in sufficient amounts in ethylene production by steam cracking and

in some refinery processes (primarily cat cracking).

Whereas in Europe, refineries satisfy on average only 20% of the chemical

industrys consumption of propene (Stam,1980), in the United States they


meet more than 40% of the consumption demand (Debreczeni,1977; Sanders

et al., 1977). Despite this high figure, refineries in the United States consume

ca. 75% of their propene production in in-house, nonchemical applications

(Spitz,1976). This propene excess is utilized in gasoline production, to

produce liquefied petroleum gas, and as heating gas.

A trend toward less severe cracking conditions and thus to increased propene

production has been observed in steam cracker plants using liquid feedstock

that have been designed since the mid-1980s. The increased consumption of

propene has also boosted the demand for processes for propene production by

catalytic dehydrogenation of propane. Various processes suitable for this

purpose have been developed and in some cases tested.

2.9.3.1. Production as Byproduct of Ethylene Production

A sour-gas-free, anhydrous C3 fraction is obtained as overhead product from

the depropanizer (a) in the processing of cracked gas. This fraction contains

the C3 hydrocarbons of the cracked gas (i.e., propane, propene, propadiene,

and propyne) as well as traces of C2 and C4 hydrocarbons. Because of its

propadiene and propyne content (depending on the cracking conditions, these

can total up to 8 mol %), this C3 fraction does not meet product specifications

and therefore requires further treatment.

Figure 2.13 shows a typical flow diagram for the further processing of the C3
fraction in a steam cracker plant. The C3 fraction is fed to a selective

hydrogenation unit (b) to remove propadiene and propyne. This

hydrogenation can be performed in the gas and liquid phases (pressure ca. 18

bar), palladium catalysts being used in both. The amount of hydrogen added

is calculated so that, on the one hand, complete conversion of C3H4 to C3H6 is

achieved, and on the other, the smallest possible amount of propene is

hydrogenated to propane. In practice a molar H2 : C3H4 ratio of ca. 1.5 has

proved suitable.

Figure 2.13. Typical flow diagram for work-up of the C3 fraction in a steam cracker plant

a) Depropanizer; b) C3 hydrogenation; c) Polymer scrubbing; d) C3 stripper; e) C3 splitter

The reaction conditions of C3 hydrogenation in the gas phase differ from those

in the liquid phase. In gas-phase hydrogenation the reaction is controlled by


means of the operating temperature, which may be between 50 and 120 C,

depending on the preparation and aging state of the catalyst.With liquid-phase

hydrogenation the reaction is controlled by the hydrogen partial pressure. The

operating temperature of 15 25 C is considerably lower in this case.

Adiabatic fixed-bed reactors with intermediate coolers, as well as isothermal

tubular reactors, have proved suitable for the exothermic hydrogenation

reaction in both phases.

Characteristic of a hydrogenation with metered addition of hydrogen is the

formation of smaller amounts of oligomers (mainly dimers and trimers). If

required by the propene product specification these can be scrubbed out

following hydrogenation with a small amount of C3 in scrubber (c). The

mixture of oligomers and scrubbing agent is recycled to the depropanizer for

C3 recovery, and the oligomers are subsequently led into the pyrolysis gasoline

fraction.

The methane introduced with the hydrogen (in liquid-phase hydrogenations,

the unconsumed hydrogen as well) is then stripped off (d) and recycled to the

cracked gas processing unit of the steam cracker to recover entrained propene.

The bottom product of the C3 stripper meets the product specification of

chemicalgrade propene.
Chemical-grade propene is purified to polymer-grade propene in a

downstream propene propane separation column (e). A single water- cooled

column designed for this purpose is illustrated in Figure 1. At this point,

reference may be made only to one advantage, which provides for the (not

essential) integration of propene propane separation in the steam cracker

plant. In modern plants the cracked gases are usually cooled to ambient

temperature in a water scrubbing tower after the oil fractionation.

The circulating water is heated to 80 to 85 C in this operation and must be

cooled again to be reused in the water scrubbing tower. The heated circulating

water can therefore be used to heat the propene propane separation stage and

thus save energy. Also, other heat consumers (e.g., the reboiler of the C3

stripper) may be supplied economically from this energy source. The propane

separated in the C3 splitter is usually (like ethane) recycled as cracking

feedstock.

2.9.3.2. Variants of Propene Propane Separation

The closely similar boiling curves of propene and propane (bp of propane

42.1 C) require highly complex separation units. The internal reflux ratio is

generally between 0.90 and 0.97 (depending on the feedstock composition,

product purities, column pressure, and number of trays).


For this reason the propene propane separation is a process unit in which the

development of energy-saving solutions is worthwhile.

Single-Column Process. In principle, separation can be carried out in a single

column (generally containing 150 200 trays), as illustrated in Figure 2.13

(e). The reflux can be condensed with coolingwater (column pressure 16 19

bar) or in air coolers (column pressure 21 26 bar). However, for the

throughputs that are common nowadays, huge column diameters are required,

and transportation of these huge units (or parts of them) by rail or road is

difficult. Thus the double column process has been developed where the

separation is carried out in two parallel columns that are smaller in diameter

and can therefore be transported more easily. This double- column process

requires only ca. 55% of the original amount of cooling water.

Double-Column Process. The double column process is illustrated in Figure

2.14. In this process, only the reflux from the second Propene separation

column (b) is condensed with cooling water. The pressure of the first column

(a) is so high (ca. 25 bar) that its overhead vapors (ca. 59 C) can be liquefied

in the reboiler of the second column (ca. 18 bar, 51 C) and serve as heat

medium. Each column produces ca. 50% of the propene product. The bottom

product from the first column is the feedstock for the second column. Heating
the first column with warm water is still possible with this process.

Figure 2.14. Propene propane separation by the doublecolumn Process a) High-


pressure C3 splitter (ca. 25 bar); b) Low-pressure C3 splitter (ca. 18 bar)

Heat Pump Process. If no cost-free (or even cost-saving) heating medium

such as the warm circulating water of the steam cracker is available, then

separation according to the heat pump process is possible (see Fig. 2.15). In

this process the higher propene propane ratio in the gas phase at lower

pressure is utilized. The overhead vapors of the C3 splitter (a), which operates

at ca. 10 bar, are heated slightly in the reflux subcooler (b) and then

compressed to a pressure that enables them to be liquefied in the reboiler (e).

Apart from the energy needed to drive the compressor (c), only a small amount

of coolingwater is required in the aftercooler (d).


Figure 2.15 . Propene propane separation by the heat pump Process. a) C3 splitter; b) Reflux
subcooler; c) Compressor; d) Postcooler; e) Condenser and reboiler

Lower Product Quality. Energy savings in the propene propane separation

can be achieved not only by appropriate process design, but also by restricting

the purity of the polymer-grade propene to the absolute minimum necessary.

Figure 4 shows the relative complexity of separation required for higher

purities, relative to a minimum propene purity of 99.0 mol %. The complexity

of separation and thus the energy consumption increases considerably with

propene purity, and even more sharply as the number of column trays

decreases.

2.9.4 Storage, Transportation, Quality Requirements


There is an extensive propene pipeline network in Texas and Louisiana in the

United States and a very limited system in Belgium in Western Europe.

In other locations propene transportation is largely by road, rail, and ship. This

discontinuous delivery and supplying of consumers requires large storage

capacities on the part of both propene producers and propene consumers.

Liquid propene is normally stored at ambient temperature in spherical

pressure tanks with diameters up to 20 m. Propene is however also stored

virtually pressureless at47 C, particularly for very large amounts. The latter

form of storage requires reliquefaction devices.

Propene is transported by road and rail in cylindrical pressurized tanks at

ambient temperature.

The standard railway tanker holds 42 t of propene. In the case of road transport

the maximum permitted overall truck weight of ca. 40 t restricts the propene

load to ca. 20 t (corresponding to 40m3). In transportation by ship, batteries

of smaller pressurized tanks as well as large atmospheric tanks with

reliquefaction devices are used.

Quality Requirements. The quality requirements placed on propene vary

widely depending on its subsequent use. Corresponding to production in steam

cracker plants, two qualities are basically employed namely, the less pure

chemical-grade propene and the highly pure polymer-grade propene.


Examples of specifications for these two product qualities are given in Table

2. Depending on different posttreatment processes (in the case of polymer-

grade propene) or different intended main uses (in the case of chemical-grade

propene) the permissible concentrations of impurities may vary within

relatively wide limits.

2.9.5 Uses of Propene

Global propene production for chemical uses increased strongly through the

1980s (WPI, 1991) equivalent to a global average annual increase of 5%.

The extent of steam cracker and refinery production of propene varies from

region to region (WPI, 1991). In the United States, refinery production has a

much greater share of total availability than in other regions due to a large

motor gasoline market that requires an adequate FCC capacity and a

preponderance of ethane feedstock for steam crackers, resulting in lower

yields of steam cracker propene. The proportion of propene produced by steam

cracking between 1980 and 1990 is given in Table 4.

The three commercial grades of propene are used for different applications.

Refinery-grade propene (i.e., 50 70% pure propene in propane) is obtained

from refinery processes. The main uses of refinery propene are in LPG for
thermal use or as an octane-enhancing component in motor gasoline.

Refinery-grade propene can also be used in some chemical syntheses

(e.g., of cumene or isopropanol).

Chemical-grade propene is used extensively for most chemical derivatives

(e.g., oxo alcohols, acrylonitrile, or polypropylene).

Polymer-grade propene contains minimal levels of impurities such as carbonyl

sulfide that can poison catalysts used in polypropylene and propylene oxide

manufacture.

2.9.5.1. Thermal Uses

Propene has a calorific value of 45 813 kJ/kg, and refinery grade propene can

be used as fuel if more valuable uses are unavailable locally (i.e., propane

propene splitting to chemicalgrade purity). Refinery-grade propene can also

be blended into LPG for commercial sale, provided no low limits exist on the

olefin content of LPG (unsaturated hydrocarbons produce more smoke than

saturated hydrocarbons and have a lower calorific value).

2.9.5.2 Motor Gasoline Uses

Propene is used as a motor gasoline component for octane enhancement, via

dimerization formation of polygasoline or alkylation.


A.) Polygasoline Production-Dimerization

Polymerization of refinery-grade propene at high temperature and pressure

with a catalyst such as phosphoric acid produces polygasoline containing

dimers, trimers, and tetramers. Polygasoline is used as a blendstock in motor

gasoline. In 1977, dimerization of propene to isohexenes at lower temperature

and pressure by using nickel complex alkylaluminum halide catalysts was

developed by Institut Francais de Petrole (Dimersol process) (Kohn, 1997).

Propene conversions of >90% are achieved. The product, dimate, has a low

vapor pressure and a high blending octane rating (R + M)/2 of 89, compared

to 86 for cat cracker gasoline and 87 for polygasoline.

B.) Alkylation

Alkylation is an acid- catalyzed reaction between propene (or other light

olefins) and isobutane to yield a highly branched higher alkane (2,3-

dimethylpentane in the case of propene) of low vapor pressure and high octane

rating(Albright, 1990) . The reaction catalyst is either hydrofluoric acid, which

gives octane ratings of 90 92, or sulfuric acid, which gives octane ratings of

88 91(Chapin et al., 1985). If hydrofluoric acid is used as catalyst a second

reaction is promoted in which hydrogen is transferred from isobutane to

propene to yield propane and isobutene. The latter then alkylates isobutane to

produce trimethylpentanes in parallel with dimethylpentane from the direct


alkylation.

The choice of catalyst depends on feedstock and process conditions as well as

octane rating.

2.9.5.3 Chemical Uses

The principal chemical uses of propene are in the manufacture of

polypropylene, acrylonitrile oxo alcohols, propylene oxide, butanal, cumene,

and propene oligomers(WPI, 1991). Polypropylene is the leading global

consumer of propene. Other uses include acrylic acid derivatives

and ethylene propene rubbers.

Polypropylene

Isotactic polypropylene is of great commercial importance. Polymerization of

chemical-grade or polymer-grade propene is carried out by using Ziegler

Natta catalysts (Natta, G. et al., 1955), in either a liquid- phase or a gas-phase

process. Polypropylene has a high melting point, good rigidity, and good

chemical and abrasion resistance. Weaknesses include poor thermal and UV

stability, but these can be improved by appropriate stabilizers. Polypropylene

is used widely, and global production exceeded 13106 t in 1990.

Acrylonitrile. Acrylonitrile is obtained by catalytic oxidation of chemical-

grade propene in the presence of ammonia (ammoxidation).

A number of commercial processes are available but the most significant is


the Sohio process (Callahan et al., 1970) in which a stoichiometric mixture of

propene, ammonia, and oxygen is reacted over a catalyst at 400 500

C.World consumption of acrylonitrile in 1990was 3.7106 t.

Propylene Oxide. Propene (polymer grade) can be converted into the ether

propylene oxide either by hydrochlorination and epoxidation, or by reaction

with an organic hydroperoxide (Stobaugh et al, 1973).

Propylene oxide is used as a raw material in the production of poly(propylene

glycols), propanolamines, and polyether polyols, which themselves are

intermediates for flexible and rigid polyurethane foams. Propylene oxide

production in 1990 was 2.75106 t, consuming some 2.2106 t of propene.

2-Propanol. 2-Propanol is produced by the hydration of chemical-grade

propene either in the presence of sulfuric acid or directly as a vapor-phase

reaction in the presence of a catalyst at 180 270 C (Stobaugh et al., 1973).

Yields >70% based on propene are achieved at high selectivity (e.g., >90 %).

2-Propanol is used extensively as a solvent in paint, cosmetics, and

pharmaceuticals. Global production of 2-propanol in 1990 was 1.4106 t,

representing 1.2106 t of propene.

Cumene. Propene can be used to alkylate benzene to produce 1-

methylethylbenzene (cumene) (Pujado, 1976). The reaction is usually


undertaken with refinery-grade propene in the presence of phosphoric acid

catalyst. Cumene is then oxidized to cumene hydroperoxide, which

decomposes to phenol and acetone. Global production of cumene was 2.3106

t in 1990, representing 0.9106t of propene.

Oligomers. Propene oligomers are liquid polymers of propene with a low

degree of polymerization. They are intermediates between polygasoline and

the high polymer polypropylenes, and contain ten or more carbon atoms per

molecule. Propene oligomers are prepared by acid- catalyzed polymerization.

Depending on the degree of polymerization the products are used as lube oil

additives or as detergent intermediates.

Hydroformylation. Propene reacts with carbon monoxide and hydrogen

(synthesis gas) at elevated temperature and pressure in the presence of a

transition-metal catalyst (e.g., a rhodium carbonyl catalyst) to give butanals

(oxo process) (Pino, et al., 1977). n-Butanal is converted into the aldol and

hydrogenated to produce 2-ethylhexanol.

Allylics. Propene can undergo many reactions to produce allylic compounds.

In these reactions a methyl hydrogen atom is substituted and the carbon

carbon double bond is preserved. The more important allylic derivatives

include allyl chloride (3- chloropropene) (Fairbairn, et al., 1947) and acrolein

(allyl aldehyde). Propene can be reacted with acetic acid in an acidic medium
to produce allyl acetate (i.e., the double bond is retained in the molecule), or

acetic acid is added to the carbon carbon double bond to yield isopropyl

acetate.

2.9.5.4 Economic Aspects

Propene is a byproduct of the production of ethylene by steam cracking or the

production of motor gasoline by catalytic cracking, so no direct manufacturing

costs are attributed to these sources. Propene demand

exceeds that from steam cracker sources, and marginal supplies are taken from

refinery sources. Propene economics are based on its value in marginal

applications: generally dimerization or alkylation for gasoline. The motor

gasoline valuation of propene is set by the octane enhancement of motor

gasoline by alkylate or Dimersol, both obtained from refinery-grade propene.

Generally, motor gasoline absorbs incremental propene supplies, and all other

uses must pay the equivalent octane value plus the costs of distilling refinery

grade to the purity required. When propene demand in motor gasoline is weak,

surplus refinery propene is returned to LPG uses or to refinery fuel. When

demand for propene is very high the value of the marginal supply rises above

the octane value to reflect the new marginal use (e.g., in polypropylene).

Propene from the dehydrogenation of propane does have a direct


manufacturing cost structure based on feedstock, utility, plant fixed costs, etc.

In an environment in which propene demand exceeds the supply from steam

crackers and catalytic crackers, propene values should cycle between a

minimum of the cash cost of propane dehydrogenation and either a level

necessary to justify the investment in further dehydrogenation plants or a

higher level set by the demand for derivatives.

2.10 Aldehyde

Aldehydes are represented by the general formula RCHO, where R can be

hydrogen or an aliphatic, aromatic, or heterocyclic group.

According to IUPAC nomenclature, aldehydes are identified by the ending

al.However, many of them still are called by their common names.

Because of their high reactivity, aliphatic aldehydes are widely used as

intermediates in organic synthesis. Sometimes the isolation of a pure aldehyde

is very difficult; in such cases, stable derivatives or oligomers are prepared

from which the aldehydes can be reisolated.


Aldehydes usually must be kept from contact with air and under certain

circumstances must be stabilized during distillation, subsequent storage, and

transportation. This applies particularly to unsaturated aldehydes, which have

a tendency to polymerize (e.g., on contact with alkali). For commerical

purposes aldehydes often are protected by the addition of stabilizers and

antioxidants and by a nitrogen atmosphere. When handling aldehydes, care

must be taken to prevent either the liquids or their vapors from coming into

contact with respiratory organs, eyes, and skin. Gloves and safety glasses are

absolutely necessary.

Aldehydes are obtained mainly via the oxo synthesis, by mild oxidation

(dehydrogenation) of primary alcohols, and by special olefin oxidation

processes. Lowconcentrations of aldehydes occur naturally in essential oils of

various plants. Acetaldehyde is an intermediate product of alcohol

fermentation; it is formed by decarboxylation of the intermediate pyruvic acid.

Aldehydes also fulfill some important biological functions, e.g., 11-cis-retinal

in the sight process or as pyridoxal in the transamination of amino acids. Their

isolation from natural substances is of commercial significance only in a few

cases, e.g., in the production of longer-chain fragrance aldehydes.

2.10.1 Physical Properties of aldehyde


Formaldehyde, the simplest saturated aliphatic aldehyde, is a gas at room

temperature. The aldehydes up to about C12 are liquids; the straightchain

aldehydes from C13 and above are solids. Because the hydrogen atom of a

formyl group has less tendency to hydrogen bond than the hydrogen atom of

a hydroxy group, the boiling points of the aldehydes are considerably lower

than those of the corresponding alcohols. The difference in boiling points

between aldehydes and alcohols with an intermediate number of carbon atoms

is 20 40 C.

Table 2.8. Physical properties of saturated aldehydes

a
at 101.3 kPa unless otherwise specified;
b
effective density at 19 C

Viscosity, density, and refractive index at 20 C increase with increasing

molecular mass. The lower homologues are mobile liquids; aldehydes from

heptanal to undecanal have an oily consistency.

There are practically no limits to the miscibility of formaldehyde and

acetaldehyde with water. However, with increasing molecular mass the

solubility of aldehydes decreases very rapidly. For example, the solubility of

hexanal in water is only 0.6wt % at 20 C. The aliphatic aldehydes are soluble

in alcohols, ethers, and other common organic solvents.

Whereas the lower aldehydes have a pungent smell, the higher homologues in

the C8 C13 range are components of nearly all perfumes and many

fragrances. Their odors become weakerwith increasing molecular mass.

The characteristic properties of most aliphatic aldehydes are their ease of

autoxidation and their tendency to trimerize and/or polymerize. Therefore,

aldehydes are protected with an inert gas atmosphere and minimal amounts of

a stabilizer are added if required (Falbe, 1983).

Because of the differences in boiling points between aldehydes and alcohols,


aldehydes prepared by dehydrogenation of the corresponding alcohols

generally are separated and purified by distillation. Apart from the lower

boiling, shortchain aldehydes, distillation is mostly carried out at reduced

pressure (Falbe, 1983). Numerous aldehydes form azeotropic mixtures with

water (see Table 2.9) and other substances.

Table 2.9 Azeotropic mixtures of aldehydes with water (Horsley, 1973)

2.10.2. Chemical Properties

The polarity of the carbonyl group of aldehydes not only facilitates the typical

aldehyde reactions addition of nucleophiles, reduction, and oxidation but

it also makes the -hydrogen atom acidic. For these reasons, aldehydes can

undergo a wide variety of reactions. The major ones are compiled in Table
2.10 (Falbe, 1983).

Addition Reactions. Because of the polarity of the carbonyl double bond,

aldehydes enter into a wide variety of nucleophilic addition reactions.

The simplest case is the intermolecular addition of one molecule of aldehyde

to another to form an aldol. These aldols (-hydroxyaldehydes) generally are

unstable and react further to form secondary products, i.e., diols, unsaturated

aldehydes, or alcohols.

Industrially, crossed aldol condensation of an aldehyde with formaldehyde is

used for the production of trimethylolpropane, pentaerythritol, and neopentyl

glycol. The aldol condensation of butanal for the production of the plasticizer

alcohol 2-ethylhexanol is of major commercial importance.


Table 2.10. General reactions of aliphatic aldehydes.
The addition of alcohols and thiols to aldehydes leads to acetals and

monothiohemiacetals; these are used for the protection of formyl groups in

synthesis because they are resistant to alkali. Reductive amination of

aldehydes with ammonia or primary amines in the presence of hydrogen or

some other reducing agent gives primary or secondary amines, respectively,

e.g.,

The reaction with secondary amines yields the corresponding enamines,

which can be hydrogenated to tertiary amines.

Classically, aldehydes were identified by the preparation of crystalline

derivatives, e.g., by reaction with hydrazine, substituted hydrazines (e.g., 2,4-

dinitrophenylhydrazine), hydroxylamine, or semicarbazide. However, now

other techniques are used.

The addition of Grignard compounds leads to secondary alcohols. The

Darzens condensation of alkyl chloroacetates in the presence of a strong base

yields aldehydes containing one additional methylene group, e.g.,


With the alkylhaloacetates in the presence of zinc, 3-hydroxycarboxylates are

obtained (Reformatsky reaction).

The addition of sodium hydrogensulfite leads to water-soluble crystalline

compounds. This reaction permits the separation of -methylbranched

aldehydes from the isomeric straightchain aldehydes (Ruhrchemie, 1974). The

reaction of hydrogen cyanide with aldehydes is of commercial significance.

The highly unstable cyanohydrins can be converted to the unsaturated nitriles

by dehydration or to the -hydroxycarboxylic acids by hydrolysis. If the

reaction of hydrogen cyanide with aldehydes is carried out in the presence of

ammonia, -amino acids are obtained.

Polymerization. The formation of cyclic, 2,4,6-trialkyl-1,3,5-trioxane type

trimers is catalyzed by various acids.

At higher temperature these compounds are unstable and revert to the

monomer. With this process pure n-alkanals can be isolated from n iso-
aldehyde mixtures.

Mainly oligomers and polymers are formed by the lower aldehydes. Polymeric

formaldehyde, referred to as paraformaldehyde, is a mixture of

polyoxymethylene glycols HO(CH2O)nH, where n is 8 100. The oligomers

of acetaldehyde, paraldehyde and metaldehyde are 2,4,6-trimethyl-1,3,5-

trioxane and the cyclic tetramer, respectively. Polyacetaldehyde is a high-

molecular mass polymer with an acetal structure.

Oxidation. A characteristic feature of most aldehydes is their great tendency

to autoxidize in a radical chain reaction to the corresponding carboxylic acids.

On an industrial scale the oxidation is usually carried out in the liquid phase

with oxygen or air. Catalysts often are added to reduce the reaction time and

lower the reaction temperature. Salts of transition metals are effective

catalysts. For special purposes, hydrogen peroxide, periodic acid, nitric acid,

potassium permanganate, chromium trioxide, and peroxo compounds are used

as oxidizing agents ((Falbe, 1983). Another special method is the melting of

aldehydes with alkali, e.g.,

Hydrogenation (Reduction). Catalytic hydrogenation leads to primary


alcohols. Both Ni and Cu catalysts have proved to be the most suitable

(Katalysatoren, 1974). Normally, the reaction is carried out in the liquid phase

on fixed-bed catalysts at 20 200 C and pressures of up to 30MPa.

Hydrogenation in the gas phase is run continuously. Because of the good heat

dissipation this is advantageous, especially with sensitive starting materials.

In addition to catalytic hydrogenation there are many other reduction

processes. Complex hydrides such as lithium aluminum hydride and sodium

borohydride are used most frequently as reducing agents.

An important reaction characteristic of aldehydes is the Meerwein Ponndorf

Verley reduction with aluminum alkoxides, in particular with aluminum

triisopropoxide. Because this reaction is reversible the carbonyl compound

formed is distilled from the alkoxide. The process is recommended for the

preparation of alcohols containing halo or nitro groups.

There are numerous other reducing agents for aldehydes; most of these are of

interest only for laboratory syntheses. A few are: potassium ammonia,

trimethyl phosphite, magnesium, aluminum mercury, trialkylboranes, and

aluminum trialkyls.

The reduction of aldehydes to the corresponding saturated hydrocarbons is of

no commercial significance. In laboratory syntheses it is achieved with


reducing agents such as hydrazine, complex metal hydrides, lithium

ammonia, or hydrogen iodide phosphorus.

2.10.3 Storage, Transportation, and Environmental Regulations

For storage and transportion, containers of stainless steel normally are used.

Vessels lined with polyethylene or other coatings are also suitable. For

aldehydes that enter the market as solutions, such as formaldehyde, glyoxal,

or glutaraldehyde, aluminum vessels or containers of standard steel should not

be used, because the acids formed by autoxidation are corrosive, and the

corrosion products can cause discoloration of the aldehyde.

Aldehydes are normally stored under a nitrogen atmosphere. Antioxidants and

stabilizers are added to prevent autoxidation. The formation of cyclic 2,4,6-

trialkyl-1,3,5-trioxane trimers is catalyzed by the presence of strong protic and

Lewis acids. The trimerization tendency of aldehydes occasionally increases

at lower temperatures, so that some compounds cannot be stored over long

periods at temperatures under 20 C without stabilization (Falbe, 1983).

Some , -unsaturated aldehydes, such as acrolein, methacrolein, and

crotonaldehyde, can react violently if handled incorrectly in the presence of

air or oxygen or some other, in particular, alkaline compound.


Because of their health hazards, saturated and unsaturated aliphatic aldehydes,

especially the lower members, should be handled only ifspecial safety

precautions are taken. Necessary measures can be found in the regulations on

dangerous materials and in the directions on accident dent prevention,

guidelines, leaflets, etc., published by the confederation of trade associations.

For the economically important aldehydes numerous transport regulations and

legal directions exist (IMO, IATA, RID, ADR, etc.). The classification of a

special aldehyde depends on its specific properties such as flash point, boiling

point, solubility in water, toxicity, and ignition temperature.

Because the aldehydes sometimes have a very unpleasant and intense odor,

special environmental measures are necessary during manufacture, storage,

and dispatch. The aldehyde containing off-gas from the production normally

is drawn off centrally and burned. The wastewaters are treated chemically and

biologically.

2.11 BUTANOLS

Butanols (butyl alcohols) are aliphatic saturated C4 alcohols (C4H9OH). There

are four structural isomers of the alcohols: two primary, one secondary, and

one tertiary; as there is an asymmetric C atom in the secondary alcohol, there


are two stereoisomers of 2-butanol.

Table 2.11 Four structural isomers of the alcohols

In addition to the systematic IUPAC nomenclature, other, nonsystematic

names for the various butanols are in common use.

1-Butanol (n-butanol) occurs in nature in compound form. It also occurs,

sometimes in high concentrations, in fusel oils obtained by fermentation.

The first industrial production of 1-butanol, around 1912, was based on the

discovery of the bacterium Clostridium acetobutylicum Weizmann, which

causes carbohydrates to ferment to give mainly acetone and 1-butanol.


The continually increasing demand for 1-butanol could only be met by

developing newmanufacturing processes, such as the hydrogenation of

crotonaldehyde formed by the aldolization of

acetaldehyde; the Reppe synthesis (propylene carbonylation); and, in

particular, the process by which most 1-butanol is produced, the

hydrogenation of n-butyraldehyde, easily obtainable by the hydroformylation

of propylene.

2-Methyl-1-propanol (isobutanol) occurs in natural products as well as in

fusel oils (74 % of the total alcohol in the product of molasses

fermentation),from which it can be isolated. The

isobutyl oil synthesis, a reaction related to methanol synthesis from CO and

H2, first made it possible to produce larger quantities of 2-methyl-1-propanol.

In the 1980s, nearly all 2-methyl-1- propanol is produced by the oxo synthesis

(propylene hydroformylation). This straight forward production process and

the attractive prices have led to a considerable expansion of the market for

isobutanol.

The secondary and tertiary butanols are obtained by hydration of the

corresponding unsaturated hydrocarbons, the butenes, whereas dehydration of

the butanol yields the corresponding butene.

Thus, 2-butanol (sec-butanol), the simplest alcohol with an asymmetric


carbon atom, can be obtained by the acid-catalyzed hydration 2 Butanols of

1- and 2-butene whereas hydration of 2- methylpropene (isobutene) yields 2-

methyl-2- propanol (tert-butanol). The ready availability of 1- and 2-butene

and 2-methylpropene from olefins derived from petroleum cracking has made

the commercial application of these reactions possible. Most 2-butanol

produced is further processed to 2-butanone (methyl ethyl ketone). The rapid

development of 2-methyl-2-propyl methyl ether as an additive for low-lead

gasolines has increased the availability of 2-methyl- 2-propanol, from which

it can be derived as byproduct.

2.11.1 PHYSICAL PROPERTIES BUTANOLS

With the exception of 2-methyl-2-propanol (mp 25.6 C) butanols are

colorless liquids. Butanols exhibit a characteristic odor; their vapors have an

irritant effect on mucous membranes and a narcotic effect in higher

concentrations.

All butanols are completely miscible with common organic solvents. Only 2-

methyl-2- propanol is completely miscible with water. The major

characteristic physical properties of butanols

are compiled in Table 2.12.

Table 2.12 Physical Properties Of Butanols


Gallant, R.W (1968), (1968) describes further physical properties of butanols.

Rizk H. A and N.Youssef (1968) as well as Rajala, G. E. and J. Crossley

(1971) have investigated the dielectric properties of butanol in detail. Davis

(1963 has set up nomograms for the densities of aqueous solutions of 1-

butanol, 2-methyl-1-propanol, and 2-methyl-2-propanol.

Further physical properties of the systems comprising water with the various

butanols have been described by several authors; Kipling (1963) reported for

1-butanol/water, Alzybejewa et al, (1964) and Schneider et al., (1966) for 2-

butanol, and Arm, (1962) for 2-methyl-2-propanol.


For the phase equilibria of the ternary systems isobutyraldehyde/2-methyl-1-

propanol isobutanol)/water and 2-butanol/2-methyl-2- propanol (tert-

butanol)/water.

Table 2.13 shows the composition of some binary azeotropic mixtures.

Table 2.13 Azeotropic mixtures (David R. Lide, 2010; Hoechst AG and

Ruhrchemie,1971;Horsley, and Tamplin,1973)

2.11.2 CHEMICAL PROPERTIES OF BUTANOLS

The chemical properties of butanols are mainly those of primary, secondary,

and tertiary alcohols. In particular the reactive primary alcohols serve as

starting materials for a wide range of reactions.

The following deserve special mention:

Dehydration: The catalytic dehydration of alcohols for the preparation of

olefins is a well-established reaction Asinger, (1971). Of the butanols 2-


methyl-2-propanol (tert-butanol) undergoes this reaction most readily. Even

with dilute sulfuric acid water is split off on warming, leading to the formation

of very pure 2- methylpropene (isobutene) (Gates et al., 1972). As 2-

methylpropene is readily obtainable from the C4-fractions of olefins from

cracked petroleum this reaction is no longer of industrial importance. The

dehydration of 2- methyl-1-propanol (isobutanol) similarly leads to the

formation of 2-methylpropene

The dehydration of 2-methyl-1-propanol to 2-methylpropene is carried out

in the presence of -Al2O3 catalyst at about 300 350 C with practically

quantitative conversion and a 2- methylpropene selectivity of over 90 %. In

an analogous manner 1-butanol can be dehydrated to a mixture of 1-butene

and cisand trans-2-butene, (Pines and Haag,1960; Knozinger and Kohne,

1966)

Selective dehydration of 1-butanol in mixtures with 2-methyl-1-propanol

can be carried out with calcium zeolite molecular sieves (Weisz,1965), due to

the difference in the spatial structure of the two butanols.

Also, at lower temperature dibutyl ether can be formed from 1-butanol in

the presence of dehydration catalysts (Knozinger and Kohne, 1966). The

branched butanols show little tendency to undergo this reaction.

Oxidation: As primary and secondary alcohols 1-butanol, 2-methyl-1-


propanol, and 2-butanol can be dehydrogenated to the corresponding carbonyl

compounds. Dehydrogenation can be carried out even at low reaction

temperatures with oxidizing agents such as manganese(IV) oxide in sulfuric

acid, nitric acid, chromic acid, or selenium dioxide (Bexten, and Weber,1975).

Even without oxidizing agents the use of suitable catalysts (e.g., Cu) at

higher temperatures enables the alcohols to be dehydrogenated to aldehyde or

ketone until the thermodynamic equilibrium has been established (Buckley

and Cox 1967; Rao et al, 1969). 2-Butanol dehydrogenates to 2-butanone

(methyl ethyl ketone) at about 350 C. (www.digitalengineeringlibrary.com)

As a tertiary alcohol 2-methyl-2-propanol cannot be selectively

dehydrogenated; splitting of the molecule occurs under extreme conditions.

Primary alcohols can be oxidized to carboxylic acids under other reaction

conditions. In addition to the standard oxidizing agents, this oxidation can be

accomplished by the reaction of the alcohol with alkali hydroxide at 275 C.

In this reaction, which is based on the observations of Guerbert, the alkali salt

of butyric acid is formed together with hydrogen and 2-ethylhexanol (Cornils

and Zilly, 1975)

The Koch-Haaf reaction can be used to prepare trimethylacetic acid from

2- methyl-1-propanol and 2-methyl-2-propanol by the addition of carbon

monoxide in the presence of sulfuric acid Koch et al., 1958).With readily


available 2-methyl-1-propanol (isobutanol) from the oxo synthesis,

trimethylacetic acid can be obtained in 89 % yield.

Alkylation: Butanols can be employed in a wide variety of alkylation

reactions. N-alkyl-, N,N-dialkyl-, or N,N,N-trialkylamines are obtainable

with ammonia and amines. The use of isomeric mixtures of alcohols leads to

the corresponding mixed amines. Ring alkylation of aromatic hydrocarbons

proceeds with butanol in the presence of Friedel-Crafts catalysts (Kozlova et

al., 1968). Esterification. Butanols can be converted in the usual manner into

butyl esters with inorganic and organic acids. The reaction is generally carried

out in the presence of acid catalysts. In certain cases excess of acid which is

to be esterified can serve as the catalyst. The rate of ester formation greatly

depends on the structure of the carboxylic acid and of the alcohol. The primary

butanols react more rapidly than 2-butanol, which in turn reacts more quickly

than 2-methyl- 2-propanol.

2.11.3 PRODUCTION OF BUTANOLS

Of the many available processes for the preparation of butanols, the

following have achieved industrial importance:

1-Butanol:

propylene hydroformylation (oxo synthesis)


Reppe synthesis crotonaldehyde hydrogenation.

2-Methyl-1-propanol:

propylene hydroformylation catalytic hydrogenation of carbon monoxide

homologization reaction

2-Butanol:

n-butene hydration

2-Methyl-2-propanol:

2-methylpropene hydration

byproduct of propylene oxide and methyl tert-butyl ether production

2.11.3.1 Oxo Synthesis.

The most important process for the manufacture of 1-butanol and 2-methyl-

1-propanol is propylene hydroformylation with subsequent hydrogenation of

the aldehydes formed.

In the oxo reaction (hydroformylation) carbon monoxide and hydrogen are

added to a carbon carbon double bond in the liquid phase in the presence of

catalysts (hydrocarbonyls or substituted hydrocarbonyls of Co, Rh, or Ru)

(Falbe, 1970). In the first reaction step aldehydes are formed with one more

C-atom than the original olefins. For olefins with more than two Catoms,

isomeric aldehyde mixtures are normally obtained. In the case of propylene


these consist of 1-butanal and 2-methylpropanal.

There are several variations of the hydroformylation process, the

differences being in the reaction conditions (pressure, temperature) as well as

the catalyst system used. The classic high-pressure process exclusively used

until the beginning of the 1970s operates at pressures of 20 30MPa (200

300 bar) CO/H2 and temperatures of 100 180 C. The catalyst is Co. It leads

to about 75 % 1-butanol and about 25 % 2-methyl-1-propanol.

The new process developments of the past few years have led to a clear

shift in the range of products. The processes operating at relatively low

pressures (15MPa , 10 50bar) use modified Rh-catalysts. The isomeric

ratios achieved are about 92 : 8 (UCC,1970) or 95 : 5 (RP-Industries,1976) 1-

butanol to 2-methyl-1-propanol. However, by the use of unmodified Rh the

percentage of 2-methyl-1- propanol can be increased to about 50 %. Catalytic

hydrogenation of the aldehydes leads to the formation of the corresponding

alcohols.

As only primary alcohols can be obtained via the oxo synthesis, it is not
possible to produce 2-butanol and 2-methyl-2-propanol by this process.

2.11.3.2 Reppe Process

1-Butanol and 2-methyl-1-propanol can also be produced on a

commercial scale by the carbonylation of propylene developed by Reppe. In

this process, developed in 1942, Reppe et al., 1953, olefins, carbon monoxide,

and water are made to react under pressure in the presence of a catalyst

(tertiary ammonium salt of polynuclear iron carbonyl hydrides). The

difference between this process and the classic Co-catalyzed

hydroformylation is that at low temperature (about 100 C) and low pressure

(0.52MPa, 5 20bar) alcohols are formed directly from the olefin.

As with the oxo synthesis the carbon monoxide can be added to both C-

atoms of the double bond which means that, when propylene is used, 1-

butanol and 2-methyl-1-propanol are obtained in a ratio of 86: 14.

The catalyst is sensitive both to air and to higher temperatures at which,

in the presence of water and CO2, it decomposes into iron carbonate. To


achieve adequate reaction rates the catalyst, carbonyl triferrate, must be

present in concentrations of >>10 % in the reaction solution; this is attained

by the presence of dissolving agents (N-alkylpyrrolidine) (Kindler and

Eisfeld,1967)

The Reppe process has not been as successful as propylene

hydroformylation with Co catalysts despite favorable n/iso ratios in the

reaction products and milder reaction conditions. This can be attributed to the

more expensive process technology.

2.11.3.3 Hydrogenation of Crotonaldehyde

Until the mid 1950s the manufacture of 1-butanol based on acetaldehyde was

the preferred process. With the development of the oxo synthesis, however, it

has lost its importance even in Japan and the USA and is no longer in use.

The individual steps of the process are as follows:

aldol condensation

2 CH3CHO CH3CH(OH)CH2CHO

splitting off of water

CH3CH(OH)CH2CHO CH3CH=CHCHO+H2O

hydrogenation

CH3CH=CHCHO+2H2 CH3CH2CH2CH2OH
Acetaldehyde is aldolized to acetaldol at normal temperature and pressure

in the presence of alkaline catalysts. With conversions of about 60 % the

acetaldol yield is about 95 %. Unreacted acetaldehyde can be distilled off and

recycled.

The removal of water and formation of aldehyde is brought about by

acidification of the acetaldol with acetic acid or phosphoric acid and

subsequent distillation, whereby the crotonaldehyde is obtained almost

quantitatively as the top product.

Various gas- and liquid-phase processes have proved their value for the

hydrogenation of crotonaldehyde to 1-butanol. Copper catalysts are

particularly useful. About 1000 kg of 1-butanol can be obtained from 1350 kg

of acetaldehyde (Kyowa, 1969).

The future economic importance of the crotonaldehyde hydrogenation

process will depend to a large extent on developments in the raw materials

market. As crude oil raw material becomes scarcer and more expensive as a

basis for the oxo synthesis, the alternative route based on ethanol from

fermentation will become increasingly competitive. The biomass substrate for

ethanol production is almost inexhaustible. The ethanol is first

dehydrogenated to form acetaldehyde, from which the synthesis can proceed.

Such a process route is an increasingly interesting alternative for tropical


countries with large supplies of cheap biomass as well as for the more

developed countries of the third world which do not have their own oil

resources (Swodenk, 1983).

2.11.3.4 Catalytic Hydrogenation of Carbon Monoxide.

In addition to the CO/H2-based methanol synthesis developed

commercially in the 1920s, other analogous processes were developed for the

manufacture of higher alcohols. Compared with the methanol synthesis,

however, these processes found only limited application. Nowadays they have

no commercial importance.

The most significant process for the production of 2-methyl-1-propanol

was the BASF isobutanol oil synthesis. Reaction products with about 50 %

methanol and 11 14 % 2-methyl-1-propanol as well as various other

components were obtained from CO and H2 at temperatures of about 430 C

and pressures of about 30MPa (300 bar) in the presence of a KOH-modified

methanol synthesis catalyst (Kyowa, 1969). BASF ceased its isobutyl oil

synthesis in 1952 when 2-methyl-1-propanol and its secondary product 2-

methylpropene could be obtained more economically using the oxo synthesis

or via a petrochemical route. Further comprehensive experiments to obtain the

desired butanol from the cheap raw materials methanol and ethanol with the
homologization reaction have so far proved fruitless. Butanols have so far

only been obtained in small amounts as byproducts.

2.11.3.5 Olefin Hydration. 2-Butanol (sec-butanol) and 2-methyl-2-propanol

(tert-butanol) can be obtained by the acid-catalyzed addition of water to 1- or

2-butene, and 2-methylpropene respectively.

The water adds at the double bond in accordance with the Markovnikov

rule. The difference in olefin reactivity can be used for the selective

manufacture of 2-butanol or 2-methyl-2- propanol from technical C4-fractions

where the two alcohols are prepared in turn and isolated. 2-Methylpropene

can be completely converted to 2-methyl-1-propanol with 65 % sulfuric acid.

Hydration of n-butene, on the other hand, requires 75 80 % H2SO4 (Kropf,

1966). Because of the tendency of the olefins to polymerize readily, the

requisite reaction temperatures must not be exceeded.

In addition to this indirect hydration another process variant is known to

industry in which the olefins are hydrated directly at elevated

temperatures(about 200 300 C) and higher pressures (10 35MPa, 100

350 bar) (Sittig, 1966)


Various catalysts have been described for direct hydration, such as

tungsten oxide and aqueous aluminum hydroxide gel suspensions. For the

hydration of 1-butene and 2-butene the best results are obtained at

temperatures between 220 and 300 C or in the case of 2- methylpropene, at

190 245 C. Direct hydration of the olefins has not yet established itself as

a serious competitor to indirect hydration (Gulf Research & Development Co.,

1958)

2.11.3.6 Other Processes.

2-Methyl-2-propanol (tert-butanol) is formed during the manufacture of

propylene oxide from 2-methylpropane in an amount of about 1.2 t per ton of

propylene oxide. In anintermediate stage tert-butyl hydroperoxide is formed

from isobutane. Another process, in which 2-methyl-2-propanol is obtained as

a by-product, is the splitting of methyl tert-butyl ether to highly pure 2-

methylpropene or 2-methyl-2-propanol. Similarly the methyl tert-butyl ether

synthesis with slight traces of water in the feedstock (methanol andC4-olefins)

produces 2-methyl-2-propanol as a byproduct (Obenaus et al., 1976)

In the literature a number of other methods for the preparation of butanol

are cited which are generally of no significance other than for laboratory

preparation.
2.11.4 QUALITY REQUIREMENTS AND CONTROL OF

BUTANOLS

On an industrial scale the four different butanols are generally marketed

as pure substances and not as isomer mixtures. Because of their many

applications both as commodities and as materials for the manufacture of

special fine chemicals a high standard of purity is demanded for the butanols.

The quality is mainly controlled by determining the physical and chemical

properties of the alcohols, such as density, refractive index, color, water

content, hydroxyl value, acid value, carbonyl value, and distillation range. The

characteristic values are normally determined according to the ASTM(1984,

DIN (1971), or the BS methods (1971).

Table 2.13 lists the standard specifications which are stipulated by the

American Society for Testing and Materials for 1-butanol, 2-butanol, and 2-

methyl-1-propanol.

Table 2.13. ASTM Specifications for butanols (ASTM, 1984)


The B. S. I. (British Standards Institution) has issued specifications for 1-

butanol (BS 508:1966) and 2-butanol (BS 1993: 1968).

The best way to verify isomer purity and detect traces of impurities is by

gaschromatographic analysis. Such analyses can be carried out rapidly and

with good reproducibility, on a routine basis, enabling the use of a high

sampling frequency.

Separation can be effected by the use of both polar and non-polar packed

columns or appropriate capillary columns.

For general purity analyses a thermal conductivity detector system is used;

trace (impurity) analysis is carried out using a flame ionization detector

system. Experience has shown that nitrogen, hydrogen, and helium are

suitable carrier gases.

The degree of purity of commercial 1-butanol, 2-methyl-1-propanol, and


2-butanol is at least 99 wt %. 2-Methyl-2-propanol is supplied both as a

product of at least 99.8 wt % purity and as an azeotropic mixture with 11 to

12 wt % water.

In addition to characteristic value determination and gas-chromatographic

analysis a number of special analyses and tests may be carried out depending

on the intended application, such as esterification tests with phthalic anhydride

and color tests with concentrated sulfuric acid or KMnO4.

2.11.5 STORAGE AND TRANSPORTATION OF BUTANOLS

Butanols can be stored or dispatched in untreated mild steel or enamelled steel

drums provided that ingress of moisture is prevented. Stainless steel

containers are also suitable. The butanols are transported in rail and road tank

cars, and in drums as well as in tanker vessels and containers. It is important

to observe the national and international safety regulations relating to aviation,

ocean-going and inland waterway shipping, rail and road transport as well as

the particular safety requirements for butanols.

Some important data relating to the safe handling, storage, and transport of

butanols are compiled in Table 2.14

Table 2.14 Safety data for butanols (Hoechst and Ruhrchemie, 1971);

Hoechst, 1975)
a
For the Federal Republic of Germany.
b
Threshold Limit Value.
c
Time-weighted average concentration.

2.11.6 USES OF BUTANOLS

1-Butanol: 1-Butanol is used principally in the field of surface coating. In

the USA, for instance, approx. 85 % of 1-butanol production finds an outlet

in this sector of the market. It is used either directly as a solvent for varnishes

or it is converted into derivatives which then serve as solvents or monomer

components.

1-Butanol cannot be used directly as a solvent in the production of

nitrocellulose lacquers; in admixture with toluene, ethanol, or various esters,

however, it is an excellent thinner.

The addition of 5 10 % of 1-butanol is also valuable in overcoming

blushing (unwanted white opacity) whichmayarise when large quantities of

thinners, particularly volatile solvents, are used.

1-Butanol is also useful for regulating the viscosity and improving the
flow properties of varnishes and for the prevention of streaking in paints

and lacquers based on spirit-soluble gums and resins.

Although not itself a solvent for substances such as polystyrene and

chlorinated rubber, 1-butanol can be successfully used in quantities up to 20%

as a diluent for the commonly used solvents for these substances, which are

mainly the esters of the saturated carboxylic acids; in particular, the acetates.

The acrylic ester of 1-butanol has become increasingly important during

the last decade. It has become an essential component of latex paints since

such coatings are not only robust but can also be produced relatively cheaply.

The 1-butyl esters of phthalic, adipic, sebacic, oleic, azelaic, stearic, and

phosphoric acids are used as plasticizers and additives for surface coatings.

The most important of these products is di-1- butyl phthalate (DBP).

Consumption is, however, stagnating and in some countries it decreased

slightly within the last years.

In the USA in particular, but also in the Federal Republic of Germany and

Belgium, 1-butanol is used for the manufacture of butylamines. 1-Butanol has

numerous applications in the plastics and textile sector. It is used as a

coagulation bath for spinning acrylic fibers and in the dyeing of poly(vinyl

alcohol) fibers.

2-Methyl-1-propanol. The fields of application of 2-methyl-1-propanol


(isobutanol) closely resemble those of 1-butanol. Indeed, because of its lower

price it can often replace 1-butanol and is accordingly used as a solvent,

diluent, and additive for nitrocellulose and synthetic resins; wetting agent;

cleaner additive; and component of printing inks and related products. Its

capacity for dissolving resins (ketone, phthalate, urea, and melamine

formaldehyde resins) is of commercial importance. As a solvent for phenol

formaldehyde resins, however, 2-methyl-1-propanol is not as effective as 1-

butanol.

Several esters exhibit excellent solvent properties; in particular, isobutyl

acetate is a very good solvent for fats, chlorinated rubber, polystyrene, and

coumarone resins. The esters of phthalic acid, adipic acid, some dicarboxylic

acids as well as phosphoric acid are used as plasticizers, particularly for PVC

and its copolymers and for cellulose derivatives.

The isobutyl esters of 2,4-dichloro- and 2,4,5- trichlorophenoxyacetic acid

exhibit herbicidal activity. For the anti-freezing properties of 2- methyl-1-

propanol in gasoline .

Also of interest is the use of 2-methyl-1-propanol as an extracting agent

in the recovery of ammonium phosphate.

2-Butanol. Practically all of the 2-butanol produced is dehydrogenated to

2-butanone (MEK or methyl ethyl ketone) which has found considerable


application because of its good solvent properties and its favorable boiling

point (79.57 C). 2-Butanol itself is also used as a solvent: in particular, when

mixed with aromatic hydrocarbons it is especially suitable as a solvent for

alkyd resins and ethyl cellulose lacquers. The ability of 2-butanol to dissolve

both water and oils is applied in the manufacture of brake fluids and cleaning

agents.

2-Butanol is converted into amines by treatment with ammonia. The

acetate is recommended as a good solvent for nitrocellulose; the xanthate is

used in ore flotation. 2-Butanol also finds application in the manufacture of

perfumes, dyestuffs, fruit essences, and wetting agents.

2-Methyl-2-propanol. Consumption of 2- methyl-2-propanol is small in

comparison with the other butanols, although its use appears to be growing.

The solvent sector is again the major field of application. 2-Methyl-2-propanol

is also used as an agent for introducing the tert-butyl group into organic

compounds (e.g., tert-butylphenol for the preparation of oil-soluble resins and

antioxidants, trinitro-tert-butyltoluene as artificial musk) and as a starting

material for the preparation of peroxides (polymerization catalysts). 2-

Methyl-2-propanol competes with 2- propanol as a gasoline additive to

prevent carburetor freezing. The methylether can be used as an antiknock

agent in gasoline. Its use as a fuel additive increased from 175 000 t in 1982
to 400 000 t in 1983.

2.11.7 ECONOMIC ASPECTS OF BUTANOLS

Both in the industrialized countries and worldwide, there is considerable

excess capacity for the manufacture of 1-butanol and 2-methyl-1- propanol,

the two major products in this group of substances (Table 2.15).

Table 2.15. Butanol capacities in the leading producer countries in 1984

Any 1-butanol production plant also can be used for the


manufacture of 2-methyl-1-propanol;
1982

Plant utilization was approx. 60 70 % in 1983. The uses can vary

considerably from one region to another. As well as direct use as a sol vent the

main areas of application are in plasticizers, butyl acetate, acrylic esters, and

glycol esters (see Table 2.16).

Table 2.16 Butanol consumption by end use, %


2.11.8 TOXICOLOGY AND OCCUPATIONAL HEALTH

The main effects of exposure to excessive concentrations of all butanols

are irritation of the mucous membranes and depression of the central nervous

system. Animal studies have shown low acute oral, dermal, and inhalation

toxicity.

There is insufficient data to make a reliable assessment of the long-term

effects in animals. The limited amount of documented experience in humans

suggests that excessive exposure can cause irritation, headache, nausea,

fatigue, and dizziness.

2.11.8.1. 1-Butanol

The lowest median lethal dose (LD50) published in the literature is 790

mg/kg (rat, oral) and 4200 mg/kg (rabbit, dermal) (Lewis and Tatken, 1984).

Inhalation of high concentrations of 1-butanol by experimental animals results


in pulmonary and eye irritation, incoordination, and narcosis. The lowest

LC50 published in the literature is 8000 ppm (rats, inhal., 4 h) (Lewis and

Tatken, 1984). Application of drops of 1-butanol into the rabbit eye produces

severe corneal irritation. 1-Butanol has a slight to moderate irritating effect on

rabbit skin (Rowe et al., 1982). 1- Butanol was found to be non-mutagenic in

the Ames test (Rowe et al., 1982).

Conjunctivitis and keratitis, sometimes accompanied by small intracorneal

vacuoles, were reported among workers exposed to 1-butanol vapors.

Intracorneal vacuoles proved to be reversible after exposure had ceased

(Browning, 1965). In a ten-year-study, workers exposed to 1-butanol at

concentrations of 200 ppm and above developed corneal inflammation

associated with a burning sensation, blurring of the vision, lacrimation, and

photophobia. The mean erythrocyte count was slightly lowered. There were

only rare complaints of irritation when the concentration was 100 ppm (Rowe

et al., 1982).

Further data indicate that eye irritation may be expected when the

concentration of 1-butanol is above 50 ppm (Browning, 1965). According to

another study humans exposed to 1-butanol showed mild irritation already at

25 ppm, followed by headaches at 50 ppm (ACGIH, 1980). Workers exposed

to 80 ppm experienced a greater hearing loss (hypoacusia) in comparison with


an unexposed group (ACGIH, 1980). Dermatitis of the fingers and hands upon

exposure to 1-butanol is also reported (Browning, 1965).

Exposure limits:

TLV (ceiling limit) 50 ppm (skin) (ACGIH, 1980),

OSHA air standard (TWA) 100 ppm (Lewis and Tatken, 1984),

MAK value 100 ppm.

2.11.8.2. 2-Butanol

The lowest LD50 published in the literature is 4400 mg/kg (rat, oral)

(ACGIH, 1980). 2-Butanol, like the other butanols, is a depressant to the

central nervous system. The lowest lethal concentration published in the

literature is 10670 ppm for mice exposed for 225 min; repeated exposure of

mice to 5330 ppm produced narcosis without death; and no acute signs of

intoxication were observed in mice exposed to 1650 ppm for 420 min (Rowe

et al., 1982).

2-Butanol caused severe corneal injury when applied to the rabbit eye but

was not found to be irritating to the rabbit skin (Rowe et al., 1982).

According to one report industrial exposure to about 100 ppm has not

resulted in any difficulties (ACGIH, 1980). Another study concludes that

irritation of the eyes, nose, and throat, headache, nausea, fatigue, and dizziness
have been experienced from excessive exposure to 2-butanol (Rowe et al.,

1982).

Exposure limits:

TLV 100 ppm (TWA); 150 ppm (STEL) (ACGIH, 1980),

OSHA air standard 150 ppm (TWA),

MAK value 100 ppm.

2.11.8.3. 2-Methyl-1-propanol

The lowest published LD50 is 2460 mg/kg (rat, oral) (Rowe et al., 1982). A

dermal LD50 of 4250 mg/kg in the rabbit indicates a low toxicity due to local

application (Rowe et al., 1982). The lowest LC50 for 2-methyl-1- propanol is

8000 ppm (rats, inhal., 4 h) (ACGIH, 1980). The compound produces narcosis

in mice that had bee intermittently exposed to 6400 ppm for a total of 136h.

The narcotic dosage is followed by slight organic changes in the liver and

kidneys. Repeated exposure of mice to 2125 ppm for 9.25 h did not cause any

adverse effect (Browning, 1965).

Application of 2-methyl-1-propanol to the rabbit eye causes moderate to

severe irritation. A moderate irritation was observed after 24-h exposure of the

rabbit skin (Lewis and Tatken, 1984).

The results of long-term studies in which a small number of rats received

the material either orally or subcutaneously suggest that 2-methyl- 1-propanol


may have a carcinogenic potential(Rowe et al., 1982). Exposure of E. coli

bacteria resulted in mutagenesis(Lewis and Tatken, 1984).

In humans, there was no evidence of eye irritation during repeated 8-h

exposure to levels in the order of 100 ppm (Rowe et al., 1982). Industrial

exposure to mixed vapors of butyl acetate and 2-methyl-1-propanol has

caused vacuolar keratitis in several workers but it is not known which of the

substances was responsible (Browning, 1965). Application of 2-methyl-1-

propanol to human skin resulted in slight erythema and hyperemia (Browning,

1965).

Exposure limits:

TLV 50 ppm (TWA), 75 ppm (STEL) (ACGIH, 1980),

OSHA air standard 100 ppm (TWA) (Lewis and Tatken, 1984),

MAK value 100 ppm.

2.11.8.4. 2-Methyl-2-propanol

The lowest LD50 is reported to be 3500 mg/kg (rat, oral) (Lewis and

Tatken, 1984). The acute effect of the badsmelling 2-methyl-2-propanol is that

of a narcotic. Prolonged contact with rabbit skin failed to cause irritation

(Rowe et al., 1982). The testing of the compound by NTP (National

Toxicology Program, U.S. Department of Health and Human Services) for


carcinogenesis bioassay began in June 1983(Lewis and Tatken, 1984).

Following application to human skin there was no reaction other than

slight erythema and hyperemia. Irritating effects, headache, nausea, fatigue,

and dizziness are reported as symptoms of excessive exposure (Rowe et al.,

1982).

Exposure limits:

TLV 100 ppm (TWA), 150 ppm (STEL) (ACGIH, 1980),

OSHA air standard 100 ppm (TWA) (Lewis and Tatken, 1984),

MAK value 100 ppm.

Das könnte Ihnen auch gefallen