Vous êtes sur la page 1sur 26

Polymerization

From Wikipedia, the free encyclopedia Jump to: navigation, search

An example of alkene polymerization, in which each Styrene monomer unit's double bond reforms as a single bond with another styrene monomer and forms polystyrene.

In polymer chemistry, polymerization is a process of reacting monomer molecules together in a chemical reaction to form three-dimensional networks or polymer chains.[1][2][3] There are many forms of polymerization and different systems exist to categorize them.

Contents
[hide]
y y y y y

1 Introduction 2 Step-growth 3 Chain-growth 4 See also 5 References

[edit] Introduction
In chemical compounds, polymerization occurs via a Homopolymers variety of reaction mechanisms that vary in complexity due to functional groups present in reacting compounds[4] and their inherent steric effects explained by VSEPR Theory. In more straightforward Copolymers polymerization, alkenes, which are relatively stable due to bonding between carbon atoms form polymers through relatively simple radical reactions; in contrast, more complex reactions such as those that involve substitution at the carbonyl group require more complex synthesis due to the way in which reacting molecules polymerize.[4]

As alkenes can be formed in somewhat straightforward reaction mechanisms, they form useful compounds such as polyethylene and polyvinyl chloride (PVC) when undergoing radical reactions,[4] which are produced in high tonnages each year[4] due to their usefulness in manufacturing processes of commercial products, such as piping, insulation and packaging. Polymers such as PVC are generally referred to as "homopolymers" as they consist of repeated long chains or structures of the same monomer unit, whereas polymers that consist of more than one molecule are referred to as copolymers (or co-polymers).[5] Other monomer units, such as formaldehyde hydrates or simple aldehydes, are able to polymerize themselves at quite low temperatures (>-80oC) to form trimers;[4] molecules consisting of 3 monomer units which can cyclize to form ring cyclic structures, or undergo further reactions to form tetramers,[4] or 4 monomer-unit compounds. Further compounds either being referred to as oligomers[4] in smaller molecules. Generally, because formaldehyde is an exceptionally reactive electrophile it allows nucleophillic addition of hemiacetal intermediates, which are generally short lived and relatively unstable "mid stage" compounds which react with other molecules present to form more stable polymeric compounds. Polymerization that is not sufficiently moderated and proceeds at a fast rate can be very hazardous. This phenomenon is known as Hazardous polymerization and can cause fires and explosions.

[edit] Step-growth
Main article: Step-growth polymerization

Step-growth polymers are defined as polymers formed by the stepwise reaction between functional groups of monomers. Most step-growth polymers are also classified as condensation polymers, but not all step-growth polymers (like polyurethanes formed from isocyanate and alcohol bifunctional monomers) release condensates, in this case we talk about addition polymers. Step-growth polymers increase in molecular weight at a very slow rate at lower conversions and reach moderately high molecular weights only at very high conversion (i.e. >95%). To alleviate inconsistencies in these naming methods, adjusted definitions for condensation and addition polymers have been developed. A condensation polymer is defined as a polymer that involves loss of small molecules during its synthesis, or contains functional groups as part of its backbone chain, or its repeat unit does not contain all the atoms present in the hypothetical monomer to which it can be degraded.

[edit] Chain-growth
Main article: Chain-growth polymerization

Chain-growth polymerization (or addition polymerization) involves the linking together of molecules incorporating double or triple chemical bonds. These unsaturated monomers (the

identical molecules that make up the polymers) have extra internal bonds that are able to break and link up with other monomers to form the repeating chain. Chain-growth polymerization is involved in the manufacture of polymers such as polyethylene, polypropylene, and polyvinyl chloride (PVC). A special case of chain-growth polymerization leads to living polymerization. In the radical polymerization of ethylene, its pi bond is broken, and the two electrons rearrange to create a new propagating center like the one that attacked it. The form this propagating center takes depends on the specific type of addition mechanism. There are several mechanisms through which this can be initiated. The free radical mechanism was one of the first methods to be used. Free radicals are very reactive atoms or molecules that have unpaired electrons. Taking the polymerization of ethylene as an example, the free radical mechanism can be divided in to three stages: chain initiation, chain propagation, and chain termination.

Polymerization of ethylene

Free radical addition polymerization of ethylene must take place at high temperatures and pressures, approximately 300C and 2000 atm. While most other free radical polymerizations do not require such extreme temperatures and pressures, they do tend to lack control. One effect of this lack of control is a high degree of branching. Also, as termination occurs randomly, when two chains collide, it is impossible to control the length of individual chains. A newer method of polymerization similar to free radical, but allowing more control involves the Ziegler-Natta catalyst, especially with respect to polymer branching. Other forms of chain growth polymerization include cationic addition polymerization and anionic addition polymerization. While not used to a large extent in industry yet due to stringent reaction conditions such as lack of water and oxygen, these methods provide ways to polymerize some monomers that cannot be polymerized by free radical methods such as polypropylene. Cationic and anionic mechanisms are also more ideally suited for living polymerizations, although free radical living polymerizations have also been developed.

[edit] See also


y y y y y y y y y y

Addition polymer Condensation polymer Cross-link Metallocene Plasma polymerization Polymer characterization Polymer physics Reversible addition fragmentation chain transfer polymerization Ring-opening polymerization Sol-gel

Ziegler-Natta catalyst

[edit] References
1. ^ Introduction to Polymers 1987 R.J. Young Chapman & Hall ISBN 0-412-22170-5 2. ^ International Union of Pure and Applied Chemistry, et al. (2000) "IUPAC Gold Book" Retrieved on 11 May 2007 from "IUPAC Gold Book" on http://goldbook.iupac.org/ 3. ^ Clayden, J., Greeves, N. et al. (2000). "Organic chemistry" Oxford 4. ^ a b c d e f g Clayden, J., Greeves, N. et al. (2000), p1450-1466 5. ^ J.M.G. Cowie "Polymers: Chemistry and Physics of Modern Materials (Chapman and Hall, 2d ed. 1991) p.4

2. polymerization, any process in which relatively small molecules, called monomers, combine chemically to produce a very large chainlike or network molecule, called a polymer. The monomer molecules may be all alike, or they may represent two, three, or more different compounds. Usually at least 100 monomer molecules must be combined to make a product that has certain unique physical propertiessuch as elasticity, high tensile strength, or the ability to form fibresthat differentiate polymers from substances composed of smaller and simpler molecules; often, many thousands of monomer units are incorporated in a single molecule of a polymer. The formation of ... (100 of 304 words)

Assorted References
y

biological aspects (in life (biology): Production of polymers)

...of polymers, long-chain molecules made of repeating units of monomers (the essential building blocks mentioned above), is a far more difficult experimental problem than the formation of monomers. Polymerization reactions tend to be dehydrations. A molecule of water is lost in the formation of a peptide from two amino acids or of a disaccharid sugar from two monomers. Dehydrating agents are...

bonding (in chemical association (chemical bonding))

...molecules into larger units held together by forces weaker than chemical bonds that bind atoms in molecules. The term is usually restricted to the formation of aggregates of like molecules or atoms. Polymerization also denotes the formation of larger units by the union of like small units but usually with chemical bonds between the smaller units.
y

borate minerals (in mineral (chemical compound): Borates)

Minerals of the borate class contain boron-oxygen groups that can link together, in a phenomenon known as polymerization, to form chains, sheets, and isolated multiple groups (see Figure 12). The silicon-oxygen (SiO4) tetrahedrons of the silicates polymerize in a manner similar to the (BO3)3- triangular groups of the borates. (For further information on such...
y

magma (in igneous rock (geology): Nature of magmas)

...silicon and oxygen is a remarkably strong one, (SiO4)4 ions are stable in magmas even at exceedingly high temperatures. They also tend to join with one another, or polymerize, to form more complex anionic groups, a tendency that is especially great in the more silicic magmas. The joining is accomplished by a sharing of oxygen ions between adjacent silicon ions...
y

synthetic rubber (in rubber (chemical compound): Polymerization methods)

Synthetic elastomers are produced on an industrial scale in either solution or emulsion polymerization methods. (Solution polymerization and emulsion polymerization are described in the article chemistry of industrial polymers.) Polymers made in solution generally have more linear molecules (that is, less branching of side chains from the main polymer chain), and they also have a narrower...
y

transition elements and their compounds (in transition element (chemical element): Transition-metal catalysts)

...processes, mostly in the petroleum and polymer (plastics, fibres) industries, in which organic molecules are isomerized, built up from simple molecules, oxidized, hydrogenated, or caused to polymerize. Only a few of the most important such processes and their catalysts can be mentioned here. Catalysts are of two physical types: homogeneous (i.e., dissolved in the reaction mixture) and...

work of Marvel (in Carl Shipp Marvel (American chemist))

...after he became a lifelong consultant for the DuPont Company in 1928. Beginning in 1933, he studied polysulfone copolymers of sulfur dioxide and ethylene, determining their structure and developing polymerization initiators. In 1937 he began to study the polymerization mechanism and structure of vinyl polymers, leading to the preparation and polymerization of new monomers. During World War II...
y

Ziegler-Natta catalyst (in Ziegler-Natta catalyst (chemistry))

any of an important class of mixtures of chemical compounds remarkable for their ability to effect the polymerization of olefins (hydrocarbons containing a double carboncarbon bond) to polymers of high molecular weights and highly ordered (stereoregular) structures.
chemical reactions (in chemical reaction: Polymerization reactions)

Polymers are high-molecular-weight compounds, fashioned by the aggregation of many smaller molecules called monomers. The plastics that have so changed society and the natural and synthetic fibres used in clothing are polymers. There are two basic ways to form polymers: (a) linking small molecules together, a type of addition reaction, and (b) combining two molecules (of the same or different...
y

combustion (in combustion (chemical reaction): Special aspects)

Formation of soot is a feature of all hydrocarbon flames. It makes the flame luminous and nontransparent. The mechanism of soot formation is accounted for by simultaneous polymerization, a process whereby molecules or molecular fragments are combined into extremely large groupings, and dehydrogenation, a process that eliminates hydrogen from molecules.
organic compounds
y

hydrocarbons (in hydrocarbon (chemical compound): Polymerization)

A single alkene molecule, called a monomer, can add to the double bond of another to give a product, called a dimer, having twice the molecular weight. In the presence of an acid catalyst, the monomer 2-methylpropene (C4H8), for example, is converted to a mixture of C8H16 alkenes (dimers) suitable for subsequent conversion to 2,2,4-trimethylpentane...

gasoline production (in gasoline (fuel))

...by catalytic cracking, the application of catalysts that facilitate chemical reactions producing more gasoline. Other methods used to improve the quality of gasoline and increase its supply include polymerization, converting gaseous olefins, such as propylene and butylene, into larger molecules in the gasoline range; alkylation, a process combining an olefin and a paraffin such as isobutane;...
o

petroleum refining (in petroleum refining: Polymerization and alkylation)

The light gaseous hydrocarbons produced by catalytic cracking are highly unsaturated and are usually converted into high-octane gasoline components in polymerization or alkylation processes. In polymerization, the light olefins propylene and butylene are induced to combine, or polymerize, into molecules of two or three times their original molecular weight. The catalysts employed consist of...
o

polyethylene (in ethylene (H2C=CH2) (chemical compound))

...carbon chains are constructed, and 2) as a starting material for other two-carbon compounds. The first of these is the single largest use of ethylene, consuming about one-half of the annual output. Polymerization (the repetitive joining of many small molecules into larger ones) of ethylene gives polyethylene, a polymer having many uses, particularly in the production of packaging films, wire...
o

polyolefins (in polyolefin (chemical compound))

any of a class of synthetic resins prepared by the polymerization of olefins. Olefins are hydrocarbons (compounds containing hydrogen [H] and carbon [C]) whose molecules contain a pair of carbon atoms linked together by a double bond. They are most often derived from natural gas or from low-molecular-weight constituents of petroleum, and their most prominent members are ethylene and propylene....
y

vinylic halides (in organohalogen compound: Reactions)

Polymerization of certain vinylic halides yields materials of economic value. Among synthetic polymers, the annual production of polyvinyl chloride, or PVC, is second only to that of polyethylene.

People The following are some people associated with "polymerization"


y

Carl Shipp Marvel (American chemist)

Other The following is a selection of items (artistic styles or groups, constructions, events, fictional characters, organizations, publications) associated with "polymerization"
y y y y y y y

catenation (chemistry) chemical reaction initiator (polymerization) plastic (chemical compound) polymer (chemistry) polymerization (chemical reaction) Ziegler-Natta catalyst (chemistry)

LINKS

Hydrogenation
From Wikipedia, the free encyclopedia Jump to: navigation, search

Catalysed hydrogenation
Process type Chemical

Food industry, petrochemical Industrial sector(s) industry, pharmaceutical industry, agricultural industry

Main technologies or sub-processes

Various transition metal catalysts, high-pressure technology

Unsaturated substrates and Feedstock hydrogen or hydrogen donors

Product(s)

Saturated hydrocarbons and derivatives

Inventor

Paul Sabatier

Year of invention

1897

Hydrogenation, to treat with hydrogen, also a form of chemical reduction, is a chemical reaction between molecular hydrogen (H2) and another compound or element, usually in the presence of a catalyst. The process is commonly employed to reduce or saturate organic compounds. Hydrogenation typically constitutes the addition of pairs of hydrogen atoms to a molecule, generally an alkene. Catalysts are required for the reaction to be usable; non-catalytic hydrogenation takes place only at very high temperatures. Hydrogen adds to double and triple bonds in hydrocarbons.[1] Because of the importance of hydrogen, many related reactions have been developed for its use. Most hydrogenations use gaseous hydrogen (H2), but some involve the alternative sources of hydrogen, not H2: these processes are called transfer hydrogenations. The reverse reaction, removal of hydrogen from a molecule, is called dehydrogenation. A reaction where bonds are broken while hydrogen is added is called hydrogenolysis, a reaction that may occur to carboncarbon and carbon-heteroatom (oxygen, nitrogen or halogen) bonds. Hydrogenation differs from protonation or hydride addition: in hydrogenation, the products have the same charge as the reactants. An illustrative example of a hydrogenation reaction is the addition of hydrogen to maleic acid to form succinic acid.[2] Numerous important applications of this petrochemical are found in pharmaceutical and food industries. Hydrogenation of unsaturated fats produces saturated fats and, in some cases, trans fats.

Contents
[hide]
y

y y

y y y

y y y

1 Process o 1.1 Substrate o 1.2 Catalysts  1.2.1 Homogeneous catalysts  1.2.2 Heterogeneous catalysts o 1.3 Hydrogen sources 2 Thermodynamics and mechanism o 2.1 Heterogeneous catalysis o 2.2 Homogeneous catalysis 3 Inorganic substrates 4 Industrial applications o 4.1 In the food industry  4.1.1 Health implications o 4.2 Hydrogenation of coal 5 History 6 Metal-free hydrogenation 7 Equipment used for hydrogenation o 7.1 Batch hydrogenation under atmospheric conditions o 7.2 Batch hydrogenation at elevated temperature and/or pressure o 7.3 Flow hydrogenation o 7.4 Industrial reactors 8 See also 9 References 10 Further reading

[edit] Process
Hydrogenation has three components, the unsaturated substrate, the hydrogen (or hydrogen source) and, invariably, a catalyst. The reaction is carried out at different temperatures and pressures depending upon the substrate and the activity of the catalyst.
[edit] Substrate

The addition of H2 to an alkene affords an alkane in the protypical reaction:


RCH=CH2 + H2 RCH2CH3 (R = alkyl, aryl)

Hydrogenation is sensitive to steric hindrance explaining the selectivity for reaction with the exocyclic double bond but not the internal double

bond. An important characteristic of alkene and alkyne hydrogenations, both the homogeneously and heterogeneously catalyzed versions, is that hydrogen addition occurs with "syn addition", with hydrogen entering from the least hindered side.[3] Typical substrates are listed in the table
Substrates for and products of hydrogenation

alkene, R2C=CR'2 alkane, R2CHCHR'2

alkyne, RCCR

alkene, cis-RHC=CHR'

aldehyde, RCHO primary alcohol, RCH2OH

ketone, R2CO

secondary alcohol, R2CHOH

ester, RCO2R'

two alcohols, RCH2OH, R'OH

imine, RR'CNR"

amine, RR'CHNHR"

amide, RC(O)NR'2 amine, RCH2NR'2

nitrile, RCN

imine, RHCNH

easily hydrogenated further

nitro, RNO2

amine, RNH2

[edit] Catalysts

With rare exceptions, no reaction below 480 C (750 K or 900 F) occurs between H2 and organic compounds in the absence of metal catalysts. The catalyst binds both the H2 and the unsaturated substrate and facilitates their union. Platinum group metals, particularly platinum, palladium, rhodium, and ruthenium, form highly active catalysts, which operate at lower temperatures and lower pressures of H2. Non-precious

metal catalysts, especially those based on nickel (such as Raney nickel and Urushibara nickel) have also been developed as economical alternatives, but they are often slower or require higher temperatures. The trade-off is activity (speed of reaction) vs. cost of the catalyst and cost of the apparatus required for use of high pressures. Notice that the Raneynickel catalysed hydrogenations require high pressures:[4][5]

Two broad families of catalysts are known - homogeneous catalysts and heterogeneous catalysts. Homogeneous catalysts dissolve in the solvent that contains the unsaturated substrate. Heterogeneous catalysts are solids that are suspended in the same solvent with the substrate or are treated with gaseous substrate.
[edit] Homogeneous catalysts

Illustrative homogeneous catalysts include the rhodium-based compound known as Wilkinson's catalyst and the iridium-based Crabtree's catalyst. An example is the hydrogenation of carvone:[6]

Hydrogenation is sensitive to steric hindrance explaining the selectivity for reaction with the exocyclic double bond but not the internal double bond. The activity and selectivity of homogeneous catalysts is adjusted by changing the ligands. For prochiral substrates, the selectivity of the catalyst can be adjusted such that one

enantiomeric product is favored. Asymmetric hydrogenation is also possible via heterogeneous catalysis on a metal that is modified by a chiral ligand.[7]
[edit] Heterogeneous catalysts

Heterogeneous catalysts for hydrogenation are more common industrially. As in homogeneous catalysts, the activity is adjusted through changes in the environment around the metal, i.e. the coordination sphere. Different faces of a crystalline heterogeneous catalyst display distinct activities, for example. Similarly, heterogeneous catalysts are affected by their supports, i.e. the material upon with the heterogeneous catalyst is bound. In many cases, highly empirical modifications involve selective "poisons". Thus, a carefully chosen catalyst can be used to hydrogenate some functional groups without affecting others, such as the hydrogenation of alkenes without touching aromatic rings, or the selective hydrogenation of alkynes to alkenes using Lindlar's catalyst. For example, when the catalyst palladium is placed on barium sulfate and then treated with quinoline, the resulting catalyst reduces alkynes only as far as alkenes. The Lindlar catalyst has been applied to the conversion of phenylacetylene to styrene.[8]

Asymmetric hydrogenation is also possible via heterogeneous catalysis on a metal that is modified by a chiral ligand.[7]
[edit] Hydrogen sources

For hydrogenation, the obvious source of hydrogen is H2 gas itself, which is typically available commercially within the storage medium of a pressurized cylinder. The hydrogenation process often uses greater than 1 atmosphere of H2, usually conveyed from the cylinders and sometimes augmented by "booster pumps". Gaseous hydrogen is produced industrially from hydrocarbons by the process known as steam reforming.[9]

Hydrogen may, in specialised applications, also be extracted ("transferred") from "hydrogen-donors" in place of H2 gas. Hydrogen donors, which often serve as solvents include hydrazine, dihydronaphthalene, dihydroanthracene, isopropanol, and formic acid.[10] In organic synthesis, transfer hydrogenation is useful for the reduction of polar unsaturated substrates, such as ketones, aldehydes, and imines.

[edit] Thermodynamics and mechanism


Hydrogenation is a strongly exothermic reaction. In the hydrogenation of vegetable oils and fatty acids, for example, the heat released is about 25 kcal per mole (105 kJ/mol), sufficient to raise the temperature of the oil by 1.6-1.7 C per iodine number drop. The mechanism of metal-catalyzed hydrogenation of alkenes and alkynes has been extensively studied.[11] First of all isotope labeling using deuterium confirms the regiochemistry of the addition:
RCH=CH2 + D2 RCHDCH2D [edit] Heterogeneous catalysis

On solids, the accepted mechanism today is called the Horiuti-Polanyi mechanism.


1. Binding of the unsaturated bond, and hydrogen dissociation into atomic hydrogen onto the catalyst 2. Addition of one atom of hydrogen; this step is reversible 3. Addition of the second atom; effectively irreversible under hydrogenating conditions.

In the second step, the metallointermediate formed is a saturated compound that can rotate and then break down, again detaching the alkene from the catalyst. Consequently, contact with a hydrogenation catalyst necessarily causes cis-trans-isomerization. This is a problem in partial hydrogenation, while in complete hydrogenation the produced trans-alkene is eventually hydrogenated.

For aromatic substrates, the first bond is hardest to hydrogenate because of the free energy penalty for breaking the aromatic system. The product of this is a cyclohexadiene, which is extremely active and cannot be isolated; in conditions reducing enough to break the aromatization, it is immediately reduced to a cyclohexene. The cyclohexene is ordinarily reduced immediately to a fully saturated cyclohexane, but special modifications to the catalysts (such as the use of the anti-solvent water on ruthenium) can preserve some of the cyclohexene, if that is a desired product.
[edit] Homogeneous catalysis

In many homogeneous hydrogenation processes,[12] the metal binds to both components to give an intermediate alkene-metal(H)2 complex. The general sequence of reactions is assumed to be as follows or a related sequence of steps:
y

binding of the hydrogen to give a dihydride complex ("oxidative addition"):

LnM + H2

LnMH2
y

binding of alkene:

LnM( 2H2) + CH2=CHR

Ln-1MH2(CH2=CHR) + L
y

transfer of one hydrogen atom from the metal to carbon (migratory insertion)

Ln-1MH2(CH2=CHR)

Ln-1M(H)(CH2-CH2R)
y

transfer of the second hydrogen atom from the metal to the alkyl group with simultaneous dissociation of the alkane ("reductive elimination")

Ln-1M(H)(CH2-CH2R)

Ln-1M + CH3-CH2R

Preceding the oxidative addition of H2 is the formation of a dihydrogen complex.

[edit] Inorganic substrates


The hydrogenation of nitrogen to give ammonia is conducted on a vast scale by the Haber-Bosch process, consuming an estimated 1% of the world's energy supply.

Dehydrogenation
From Wikipedia, the free encyclopedia

Jump to: navigation, search Dehydrogenation is a chemical reaction that involves the elimination of hydrogen (H2). It is the reverse process of hydrogenation. Dehydrogenation reactions may be either large scale industrial processes or smaller scale laboratory procedures. There are a variety of classes of dehydrogenations:
y

y y

Aromatization - Six-membered alicyclic rings can be aromatized in the presence of hydrogenation catalysts, the elements sulfur and selenium, or quinones (such as DDQ). Oxidation - The conversion of alcohols to ketones or aldehydes can be effected by metal catalysts such as copper chromite. In the Oppenauer oxidation, hydrogen is transferred from one alcohol to another to bring about the oxidation. Dehydrogenation of amines - amines can be converted to nitriles using a variety of reagents, such as IF5. Dehydrogenation of paraffins and olefins - paraffins like n-pentane and isopentane can be converted to pentene and isoprene using chromium (III) oxide as a catalyst at 500 degree C.

Dehydrogenation converts saturated fats to unsaturated fats. Enzymes that catalyze dehydrogenation are called dehydrogenases.

[edit] References

Alkylation
From Wikipedia, the free encyclopedia Jump to: navigation, search "Alkylating agent" redirects here. For the class of drugs, see alkylating antineoplastic agent.

Alkylation is the transfer of an alkyl group from one molecule to another. The alkyl group may be transferred as an alkyl carbocation, a free radical, a carbanion or a carbene (or their equivalents).[1] Alkylating agents are widely used in chemistry because the alkyl group is probably the most common group encountered in organic molecules. Many biological target molecules or their synthetic precursors are composed of an alkyl chain with specific functional groups in a specific order. Selective alkylation, or adding parts to the chain with the desired functional groups, is used, especially if there is no commonly available biological precursor. Alkylation with only one carbon is termed methylation. In oil refining contexts, alkylation refers to a particular alkylation of isobutane with olefins. It is a major aspect of the upgrading of petroleum.[2] In medicine, alkylation of DNA is used in chemotherapy to damage the DNA of cancer cells. Alkylation is accomplished with the class of drugs called alkylating antineoplastic agents.

Benzene Friedel-Crafts alkylation.

Contents
[hide]
y

y y y y

1 Alkylating agents o 1.1 Nucleophilic alkylating agents o 1.2 Electrophilic alkylating agents o 1.3 Carbene alkylating agents 2 In biology 3 Oil refining 4 References 5 External links

[edit] Alkylating agents


Alkylating agents are classified according to their nucleophilic or electrophilic character.
[edit] Nucleophilic alkylating agents

Nucleophilic alkylating agents deliver the equivalent of an alkyl anion (carbanion). Examples include the use of organometallic compounds such as Grignard (organomagnesium), organolithium, organocopper, and organosodium reagents. These compounds typically can add to an electrondeficient carbon atom such as at a carbonyl group. Nucleophilic alkylating agents can also displace halide substituents on a carbon atom. In the presence of catalysts, they also alkylate alkyl and aryl halides, as exemplified by Suzuki couplings.
[edit] Electrophilic alkylating agents

Electrophilic alkylating agents deliver the equivalent of an alkyl cation. Examples include the use of alkyl halides with a Lewis acid catalyst to alkylate aromatic substrates in Friedel-Crafts reactions. Alkyl halides can also react directly with amines to form C-N bonds; the same holds true for other nucleophiles such as alcohols, carboxylic acids, thiols, etc. Electrophilic, soluble alkylating agents are often very toxic, due to their ability to alkylate DNA. They should be handled with proper PPE. This mechanism of toxicity is also responsible for the ability of some alkylating agents to perform as anti-cancer drugs in the form of alkylating antineoplastic agents, and also as chemical weapons such as mustard gas. Alkylated DNA either does not coil or uncoil properly, or cannot be processed by information-decoding enzymes. This results in cytotoxicity with the effects of

inhibition the growth of the cell, initiation of programmed cell death or apoptosis. However, mutations are also triggered, including carcinogenic mutations, explaining the higher incidence of cancer after exposure. Alcohols and phenols can be alkylated to give alkyl ethers:
R-OH + R'-X R-O-R' + H-X

The produced acid HX is removed with a base, or, alternatively, the alcohol is deprotonated first to give an alkoxide or phenoxide. For example, dimethyl sulfate alkylates the sodium salt of phenol to give anisole, the methyl ether of phenol. The dimethyl sulfate is dealkylated to sodium methylsulfate.[3]
Ph-O Na+ + Me2SO4 Ph-O-Me + Na+ MeSO4

On the contrary, the alkylation of amines introduces the problem that the alkylation of an amine makes it more nucleophilic. Thus, when an electrophilic alkylating agent is introduced to a primary amine, it will preferentially alkylate all the way to a quaternary ammonium cation.
R-NH2 R-NH-R' R-N(R')2 R-N(R')3+ (alkylating agent omitted for clarity)

If the quaternary ammonium is not the desired product, more circuitious routes such as reductive amination are necessary.
[edit] Carbene alkylating agents

Carbenes are extremely reactive and are known to attack even unactivated C-H bonds. Carbenes can be generated by elimination of a diazo group. A metal can form a carbene equivalent called a transition metal carbene complex.

[edit] In biology
Main article: methylation

Methylation is the most common type of alkylation, being associated with the transfer of a methyl group. Methylation is distinct from alkylation in that it is specifically the transfer of one carbon, whereas alkylation can refer to the transfer of long chain carbon groups. Methylation in nature is typically effected by vitamin B12-derived enzymes, where the methyl group is carried by cobalt. In methanogenesis, coenzyme M is methylated by tetrahydromethanopterin.

Electrophilic compounds may alkylate different nucleophiles in the body. The toxicity, carcinogenity, and paradoxically, cancer cell-killing abilities of different DNA alkylating agents are an example.

[edit] Oil refining


In a standard oil refinery process, isobutane is alkylated with lowmolecular-weight alkenes (primarily a mixture of propene and butene) in the presence of a strong acid catalyst, either sulfuric acid or hydrofluoric acid. In an oil refinery it is referred to as a sulfuric acid alkylation unit (SAAU) or a hydrofluoric alkylation unit, (HFAU). Refinery workers may simply refer to it as the alky or alky unit. The catalyst protonates the alkenes (propene, butene) to produce reactive carbocations, which alkylate isobutane. The reaction is carried out at mild temperatures (0 and 30 C) in a two-phase reaction. Because the reaction is exothermic, cooling is needed: SAAU plants require lower temperatures so the cooling medium needs to be chilled, for HFAU normal refinery cooling water will suffice. It is important to keep a high ratio of isobutane to alkene at the point of reaction to prevent side reactions which produces a lower octane product, so the plants have a high recycle of isobutane back to feed. The phases separate spontaneously, so the acid phase is vigorously mixed with the hydrocarbon phase to create sufficient contact surface. The product is called alkylate and is composed of a mixture of high-octane, branched-chain paraffinic hydrocarbons (mostly isopentane and isooctane). Alkylate is a premium gasoline blending stock because it has exceptional antiknock properties and is clean burning. Alkylate is also a key component of avgas. The octane number of the alkylate depends mainly upon the kind of alkenes used and upon operating conditions. For example, isooctane results from combining butylene with isobutane and has an octane rating of 100 by definition. There are other products in the alkylate, so the octane rating will vary accordingly. Since crude oil generally contains only 10 to 40 percent of hydrocarbon constituents in the gasoline range, refineries use a fluid catalytic cracking process to convert high molecular weight hydrocarbons into smaller and more volatile compounds, which are then converted into liquid gasoline-size hydrocarbons. Alkylation processes transform low molecular-weight alkenes and iso-paraffin molecules into larger iso-paraffins with a high octane number.

Combining cracking, polymerization, and alkylation can result in a gasoline yield representing 70 percent of the starting crude oil. More advanced processes, such as cyclicization of paraffins and dehydrogenation of naphthenes forming aromatic hydrocarbons in a catalytic reformer, have also been developed to increase the octane rating of gasoline. Modern refinery operation can be shifted to produce almost any fuel type with specified performance criteria from a single crude feedstock. In the entire range of refinery processes, alkylation is a very important process that enhances the yield of high-octane gasoline. However, not all refineries have an alkylation plant. The oil and gas journal annual survey of worldwide refining capacities for January 2007 lists many countries with no alkylation plants at their refineries. Refineries examine whether it makes sense economically to install alkylation units. Alkylation units are complex, with substantial economy of scale. In addition to a suitable quantity of feedstock, the price spread between the value of alkylate product and alternate feedstock disposition value must be large enough to justify the installation. Alternative outlets for refinery alklylation feedstocks include sales as LPG, blending of C4 streams directly into gasoline and feedstocks for chemical plants. Local market conditions vary widely between plants. Variation in the RVP specification for gasoline between countries and between seasons dramatically impacts the amount of butane streams that can be blended directly into gasoline. The transportation of specific types of LPG streams can be expensive so local disparities in economic conditions are often not fully mitigated by cross market movements of alkylation feedstocks. The availability of a suitable catalyst is also an important factor in deciding whether to build an alkylation plant. If sulfuric acid is used, significant volumes are needed. Access to a suitable plant is required for the supply of fresh acid and the disposition of spent acid. If a sulfuric acid plant must be constructed specifically to support an alkylation unit, such construction will have a significant impact on both the initial requirements for capital and ongoing costs of operation. Alternatively it is possible to install a WSA Process unit to regenerate the spent acid. No drying of the gas takes place. This means that there will be no loss of acid, no acidic waste material and no heat is lost in process gas reheating. The selective condensation in the WSA condenser ensures that the regenerated fresh acid will be 98% w/w even with the humid process gas. It is possible to combine spent acid regeneration with

disposal of hydrogen sulfide by using the hydrogen sulfide as a fuel.[4] The second main catalyst option is hydrofluoric acid. Rates of consumption for HF acid in alkylation plants are much lower than for sulfuric acid. HF acid plants can process a wider range of feedstock mix with propylenes and butylenes. HF plants also produce alkylate with better octane rating than sulfuric plants. However, due to the hazardous nature of the material, HF acid is produced at very few locations and transportation must be managed

alkylation, in petroleum refining, chemical process in which light, gaseous hydrocarbons are combined to produce high-octane components of gasoline. The light hydrocarbons consist of olefins such as propylene and butylene and isoparaffins such as isobutane. These compounds are fed into a reactor, where, under the influence of a sulfuric-acid or hydrofluoric-acid catalyst, they combine to form a mixture of heavier hydrocarbons. The liquid fraction of this mixture, known as alkylate, consists mainly of isooctane, a compound that lends excellent antiknock characteristics to unleaded gasolines.

CRACKING ALKANES
This page describes what cracking is, and the differences between catalytic cracking and thermal cracking used in the petrochemical industry.

Cracking
What is cracking? Cracking is the name given to breaking up large hydrocarbon molecules into smaller and more useful bits. This is achieved by using high pressures and temperatures without a catalyst, or lower temperatures and pressures in the presence of a catalyst.

The source of the large hydrocarbon molecules is often the naphtha fraction or the gas oil fraction from the fractional distillation of crude oil (petroleum). These fractions are obtained from the distillation process as liquids, but are re-vaporised before cracking. There isn't any single unique reaction happening in the cracker. The hydrocarbon molecules are broken up in a fairly random way to produce mixtures of smaller hydrocarbons, some of which have carbon-carbon double bonds. One possible reaction involving the hydrocarbon C15H32 might be:

Or, showing more clearly what happens to the various atoms and bonds:

This is only one way in which this particular molecule might break up. The ethene and propene are important materials for making plastics or producing other organic chemicals. The octane is one of the molecules found in petrol (gasoline).

Catalytic cracking Modern cracking uses zeolites as the catalyst. These are complex aluminosilicates, and are large lattices of aluminium, silicon and oxygen atoms carrying a negative charge. They are, of course, associated with positive ions such as sodium ions. You may have come across a zeolite if you know about ion exchange resins used in water softeners. The alkane is brought into contact with the catalyst at a temperature of

about 500C and moderately low pressures. The zeolites used in catalytic cracking are chosen to give high percentages of hydrocarbons with between 5 and 10 carbon atoms particularly useful for petrol (gasoline). It also produces high proportions of branched alkanes and aromatic hydrocarbons like benzene. For UK A level (and equivalent) purposes, you aren't expected to know how the catalyst works, but you may be expected to know that it involves an ionic intermediate.
Note: You should check your syllabus to find out exactly what you need to know. If you are studying a UK-based syllabus and haven't got one, follow this link. Use the BACK button on your browser to return quickly to this page.

The zeolite catalyst has sites which can remove a hydrogen from an alkane together with the two electrons which bound it to the carbon. That leaves the carbon atom with a positive charge. Ions like this are called carbonium ions (or carbocations). Reorganisation of these leads to the various products of the reaction.

Note: If you are interested in other examples of catalysis in the petrochemical industry, you should follow this link. It will lead you to information on reforming and isomerisation (as well as a repeat of what you have just read about catalytic cracking). Use the BACK button on your browser if you want to return quickly to this page.

Thermal cracking In thermal cracking, high temperatures (typically in the range of 450C to 750C) and pressures (up to about 70 atmospheres) are used to break the large hydrocarbons into smaller ones. Thermal cracking gives mixtures of products containing high proportions of hydrocarbons with double bonds - alkenes.
Warning! This is a gross oversimplification, and is written to satisfy the needs of one of the UK A level Exam Boards (AQA). In fact, there are several versions of thermal cracking designed to produce different mixtures of products. These use completely different sets of conditions. If you need to know about thermal cracking in detail, a Google search on thermal cracking will throw up lots of useful leads. Be careful to go to industry (or similarly reliable) sources.

Thermal cracking doesn't go via ionic intermediates like catalytic cracking. Instead, carbon-carbon bonds are broken so that each carbon atom ends up with a single electron. In other words, free radicals are formed.

Reactions of the free radicals lead to the various products.

Where would you like to go now? To the alkanes menu . . . To the menu of other organic compounds . . . To Main Menu . . .

Jim Clark 2003

Vous aimerez peut-être aussi