Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Isomerization

An intramolecular isomerisation that involves the breaking or making of bonds is a


special case of a molecular rearrangement.

From: Comprehensive Biotechnology (Third Edition), 2019

Related terms:

Glucose, Selectivity, Dehydration, Polymerization, Ligand, Zeolite, Alkene, Enzyme,


Isomer, Oxidation

View all Topics

Oil Refining and Products


Abdullah M. Aitani, in Encyclopedia of Energy, 2004

3.2 Isomerization
Isomerization is an intermediate, fed preparation-type process. There are more than
200 units worldwide, with a processing capacity of 1.5 million barrels/day of light
paraffins. Two types of units exist: C4 isomerization and C5/C6 isomerization. A C4
unit will convert normal butane into isobutane, to provide additional feedstock for
alkylation units, whereas a C5/C6 unit will isomerize mixtures of C5/C6 paraffins,
saturate benzene, and remove naphtenes.

Isomerization is similar to catalytic reforming in that the hydrocarbon molecules are


rearranged, but unlike catalytic reforming, isomerization just converts normal paraf-
fins to isoparaffins. The greater value of branched paraffins over straight paraffins is a
result of their higher octane contribution. The formation of isobutane is a necessary
step to produce alkylate gasoline or methyl tertiary butyl ether (MTBE). The extent
of paraffin isomerization is limited by a temperature-dependent thermodynamic
equilibrium. For these reactions, a more active catalyst permits a lower reaction
temperature and that leads to higher equilibrium levels. Isomerization of paraffins
takes place under medium pressure (typically 30 bar) in a hydrogen atmosphere.

C4 isomerization produces isobutane feedstock for alkylation. Platinum or another


metal catalyst, alumina chloride, is used for the higher temperature processes. In a
typical low-temperature process where only aluminum chloride is used, the feed to
the isomerization unit is n-butane or mixed butanes combined with hydrogen (to
inhibit olefin formation). C5/C6 isomerization increases the octane number of the
light gasoline components n-pentane and n-hexane, which are found in abundance
in straight-run gasoline. The basic C5/C6 isomerization process is essentially the
same as butane isomerization.

> Read full chapter

Superacids
George A. Olah, G.K. Surya Prakash, in Encyclopedia of Physical Science and Tech-
nology (Third Edition), 2003

III.E Isomerization
The isomerization of hydrocarbons is of practical importance. Isomeric dialkyl-
benzenes, such as xylenes, are starting materials for plastics and other products.
Generally, the need is for only one of the possible isomers, and thus there is a po-
tential for intraconversion (isomerization). Straight-chain alkanes with five to eight
carbon atoms have considerably lower octane numbers than their branched isomers,
and hence there is a need for higher-octane branched isomers. Isomerizations
are generally carried out under thermodynamically controlled conditions and lead
to equilibria. The ionic equilibria in superacid systems generally favor increasing
amounts of the higher-octane branched isomers at lower temperatures.

Lewis-acid-catalyzed isomerization of alkanes can be effected with various systems.


Superacid-catalyzed reactions can be carried out at much lower temperatures, even
at or below room temperature, and thus provide more of the branched isomers. This
is of particular importance in preparing lead-free gasoline. Increasing the octane
number by this means is preferable to the addition of higher-octane aromatics or
olefins, which may pose environmental or health-hazard problems.

Of the many important superacid-catalyzed isomerizations, the isomerization tricy-


clo [5.2.1.02,6] decane to adamantane is unique, a reaction discovered by Schleyer.

Isomerization of alkylaromatics can also be effectively carried out with superacids.

> Read full chapter


Automotive Gasoline
Scott A. Stout, ... Allen D. Uhler, in Environmental Forensics, 1964

18.3.5 Isomerization
Isomerization of low boiling (<C6) normal hydrocarbons to saturated branched
hydrocarbons was another refining development triggered by the need for octane,
that was introduced in 1943 (Speight, 1991). In this process various catalysts and
reactor conditions are used to convert C4 to C6 feed streams (butane, pentane, and/or
pentane-hexane) into their various isomeric equivalents, i.e., isobutane, isopentane,
2,2-dimethlybutane, 2,3-dimethylbutane, 2-methylpentane, and 3-methylpentane.
The specific reaction conditions during isomerization (type and age of catalyst, feed
stream and rate, and temperature) will yield rather specific distributions of isomers,
collectively known as isomerate.

Because isomerate is a common blending component in gasoline the distribution


among these C4–C6 isomers can be used to distinguish among gasolines containing
different isomerate blends. For example, it has been shown that the ratio between
isopentane (i-C5) and pentane (n-C5) or between 2-methylpentane and 3-methylpen-
tane can vary significantly depending upon the isomerization reaction conditions
(Pines, 1981). For example, the i-C5/(i-C5 + n-C5) ratio of LSRG (~0.40–0.50; see
Figure 18.3.1) is significantly increased following isomerization, being greatest in
the presence of pressurized hydrogen (~0.60–0.70; Pines, 1981; Gary and Handwerk,
1984). The degree of blending between LSRG and isomerate will largely determine
the i-C5/(i-C5+ n-C5) ratio in the finished gasoline (Beall et al., 2002). Therefore, this
ratio can be useful in environmental forensic investigations that require distinction
between fugitive gasolines in the environment. Of course, the high volatility of these
<C6 compounds warrants caution due to the potential effects of evaporation on
environmental samples.

> Read full chapter

Gasoline Manufacturing Processes


Surinder Parkash, in Refining Processes Handbook, 2003

HYDROCARBON CONTAMINANTS
The tendency of the catalyst to coke or sludge is minimal. The process therefore
offers great flexibility with respect to the amount of hydrocarbons other than C5/C6.
Sharp fractionation is not required to prevent C6 cyclics and C7 from entering the
isomerization reactor. The effect of some of these hydrocarbons follows.

Olefins
The isomerization catalyst can tolerate up to 2% of C5/C6 olefins. Therefore, the feed
from an FCCU or thermal cracker cannot be handled in an isomerization unit. Large
quantities of olefins, if present in the feed, physically coat the catalyst following
polymerization.

Cyclic Compounds
Cyclic compounds, if present in the feed, are adsorbed on the catalyst, reducing
the active sites available for paraffin isomerization. Therefore, if the feed contains
significant amounts of cyclic compounds, such as benzene, the catalyst inventory in
the reactor has to be increased. Unsaturated cyclic hydrocarbons consume consider-
able quantities of hydrogen, resulting in exothermic reactions, which is undesirable
from an isomerization equilibrium view point. Benzene is rapidly hydrogenated
and converted to cyclohexanes. Cyclohexanes and other C6 naphthenes are partially
converted to C6 paraffins.

C7 Hydrocarbons
C7 hydrocarbons crack readily to C3 and C4, and those that do not hydrocrack are
isomerized to a mixture having a lower octane number than C5 or C6. C7 naphthenes
have an effect similar to C6 naphthenes.

> Read full chapter

Applications of Coordination Chem-


istry
MD.K. Nazeeruddin, M. Grätzel, in Comprehensive Coordination Chemistry II, 2003

9.16.3.5 MLCT Transitions in Geometrical Isomers


Isomerization is another approach for tuning the spectral properties of metal com-
plexes.57–59 The UV–vis absorption spectrum of the trans-dichloro complex (35) in
DMF solution shows at least three MLCT absorption bands in the visible region at
690, 592, and 440 nm. Alternatively, the cis-dichloro complex (33) in DMF solution
shows only two distinct broadbands in the visible region at 590 and 434 nm,
assigned as MLCT transitions. The lowest energy MLCT band in the trans-complexes
((35)–(37)) is significantly red shifted compared to that in the corresponding cis-
complexes ((22), (33), (34)) (Figure 9). This red shift is due to stabilization of the
LUMO of the dcbpy ligand in the trans species relative to the cis species. The red
shift (108 nm) of the lowest energy MLCT absorption of the trans-dichloro complex
(35) compared that of the trans-dithiocyanato complex (37) is due to the strong
-donor property of the Cl− ligand compared to NCS−. The chloride ligands cause
destabilization of the metal t2g orbitals, raising them in energy closer to the ligand
π* orbitals, resulting in lower energy MLCT transitions.

Figure 9. UV–visible absorption spectra of the complexes (22) and (37) in ethanol
solution at room temperature.
> Read full chapter

Polymer Reactions
Ernest Maréchal, in Comprehensive Polymer Science and Supplements, 1989

1.3.1.2(ii) Other Constitutional Rearrangements


Many examples are known such as the Fries rearrangement in aromatic polyesters
or in some poly(methyl methacrylates) with poly(spiropyran) groups.33

Tirrell et al.57–60 studied the arrangement of poly(chloromethylthiiranes) (PCMT) and


of poly-(3-chlorothietanes) (Scheme 3).

Scheme 3.

The 13C NMR spectrum of a sample of PCMT stored at room temperature for three
months shows that the reaction in Scheme 4 takes place.

Scheme 4.

The repeating unit isomerization of PCMT occurs in the bulk and in non-nucle-
ophilic solvents such as chloroform, dichloromethane and nitrobenzene. Regardless
of the medium, the rearrangement stops when about 40% of the initial PCMT
monomer units is isomerized. The rate-determining step is the attack of sulfur atom
on the carbon–chloride bond.

The separation of the reactive groups (S and CCl) by as many as five bonds does not
preclude the isomerization from taking place, as found for poly[(4-chlorobutyl)thi-
irane] (1), this is consistent with the chemistry of low molecular weight chloroalkyl-
sulfides.

Tirrell et al.60 showed that poly(epichlorohydrin) does not undergo repeating unit
isomerization; however substituted polyether carrying functional groups placed five
bonds away from the backbone heteroatom could undergo isomerization due to
anchimeric assistance.

> Read full chapter

Catalysis, Homogeneous
Piet W.N.M. van Leeuwen, in Encyclopedia of Physical Science and Technology (Third
Edition), 2003

III.B.3 Asymmetric Isomerization


An important application of an isomerization is found in the Takasago process for the
commercial production of (−)menthol from myrcene. The catalyst used is a rhodium
complex of BINAP. The BINAP complex is an asymmetric ligand based on the
atropisomerism of substituted dinaphthyl (Fig. 33). It was first introduced by Noyori.
Atropisomers of diphenyl and the like are formed when ortho substituents do not
allow rotation around the central carbon-carbon bond. As a result two enantiomers
are formed.

FIGURE 33. The two enantiomers of BINAP.

For asymmetric hydrogenation, transfer hydrogenation, and isomerization of double


bonds using both ruthenium and rhodium complexes BINAP has been extensively
used. The synthesis of menthol is given in the reaction scheme, Fig. 34. The key
reaction is the enantioselective isomerization of the allylamine to the asymmetric
enamine. It is proposed that this reaction proceeds via an allylic intermediate.
FIGURE 34. The Takasago process for (–)menthol.

This is the only step that needs to be steered to the correct enantiomer, since the
other two are produced in the desired stereochemistry with the route depicted.
Of the eight possible isomers only this one (1R,3R,4S) is important. After the
enantioselective isomerization the enamine is hydrolyzed. A Lewis acid catalyzed
ring closure gives the menthol skeleton. In a subsequent step the isopropenyl
group is hydrogenated over a heterogeneous Raney nickel catalyst. Asymmetric
catalysis involving metal-catalyzed hydrogenations and isomerization is becoming
increasingly important in the production of pharmaceuticals, agrochemicals, and
flavors and fragrances.

> Read full chapter

Ring-Opening Polymerization and Spe-


cial Polymerization Processes
S. Kobayashi, in Polymer Science: A Comprehensive Reference, 2012

4.15.2.2.2 Double isomerization polymerization


All the above-mentioned CROPs involve isomerization one time during the polymer-
ization (single isomerization polymerization (SIP)). However, the equilibrium in the
propagating species between an oxazolinium halide and an N-haloethylamide brings
an interesting yet complicated feature in the polymerization of OZO derivatives. A
pseudo-urea is a compound having an –N=C(NR2)–O– functional group, which is an
isomeric form of urea. Then N,N-dialkylamino-2-oxazolines undergo CROP which
was expected to give a pseudo-polyurea. However, the cationic polymerization of
N,N-dialkylamino-2-oxazolines showed a unique polymerization behavior.
The cationic polymerization of 2-(1-pyrrolidinyl)-2-oxazoline (PyOZO) gave two
different polymers by the selection of initiator.18 With methyl trifluoromethane-
sulfonate (triflate) (MeOTf ) or methyl p-toluenesulfonate (tosylate) (MeOTs), the
polymerization of PyOZO yielded poly(N-(1-pyrrolidinecarbonylimino)ethylene) ac-
cording to the usual CROP (SIP). On the other hand, the polymerization of PyOZO
with RX gave poly((1,3-diazolidin-2-one-1,3-diyl)tetramethylene) via DIP (Scheme
11).19 In the latter polymerization, the OZO ring was opened and rearranged to a
five-membered cyclic urea unit.

Scheme 11. Normal polymerization and double isomerization polymerization (DIP)


of PyOZO monomer.

The polymerization mechanism of the SIP and DIP is explained as follows: In the
initiation, N-alkylated oxazolinium salt is first formed. With the sulfonate initiator,
the propagation via the oxazolinium species occurs to induce the SIP process, as the
nucleophilicity of the counteranion is weak. When the counteranion of the salt is suf-
ficiently nucleophilic as in the case of halides (X = Br, Cl), it catalyzes the rearrange-
ment of the oxazolinium ion to 3-methyl-1-azonia-3-azaspiro[4.4]nonan-2-one salt,
a spiro-structure, via a covalent-type alkyl halide species as an intermediate. This
isomerization to form the spiro-intermediate is the key step of the DIP. Its formation
is preferred since it is more thermodynamically stable than the oxazolinium ion, due
to the more stable C=O bond than the C=N bond (Scheme 12). This spiro salt is
sufficiently electrophilic to suffer the attack of the counteranion or the monomer.
The attack of the counteranion exclusively occurs at the pyrrolidinium ring, and a
covalent ethyleneurea species is generated selectively.20

Scheme 12. Mechanism proposed for DIP of PyOZO monomer.


The DIP has been extended to a bifunctional monomer, 1,4-bis(2-oxa-
zolin-2-yl)piperazine (BOP), which yielded a linear polyurea, poly((1,3-diazo-
lidin-2-one-1,3-diyl)ethylene) (PBOP).21 When BOP was heated to 100 °C or above
with methyl iodide or benzyl bromide in benzonitrile, the DIP proceeded and PBOP
was obtained as brown needles whose melting point is 210 °C (Scheme 13).

Scheme 13. DIP of a bis-oxazoline monomer of BOP.

It is very unusual that the polymerization of the bifunctional monomer of cyclic rigid
structure gives linear polymer. The polymerization mechanism for DIP of BOP can
be explained as shown in Scheme 14.

Scheme 14. Mechanism proposed for DIP of BOP monomer.

The polymerization of BOP with 2 mol% of MeOTf in benzonitrile at 150 °C, on


the other hand, quantitatively produced an insoluble gel-like product after 100 h,
whose IR spectrum showed the presence of urea and oxazoline groups. Obviously,
two oxazoline moieties in BOP polymerized independently, which results in the
branching and cross-linking of polymer.

A polymer having a high solubility was prepared from the SIP of


2-(1,3,3-trimethyl-6-azabicyclo[3.2.1]oct-6-yl)-2-oxazoline (TAOZO), although its
DIP did not proceed. This polymer was soluble in water as well as in common
organic solvents.22 The introduction of o-phenylene ring into the polymer main
chain strongly reduced its solubility. The DIP of 2-(2-isoindolinyl)-2-oxazoline (IIO-
ZO) with MeI produced an ethyleneurea-type polymer, which was partly soluble in
chloroform.23

> Read full chapter


Engineering Perspectives in Biotech-
nology
Chrysanthi Pateraki, ... Vasiliki Kachrimanidou, in Comprehensive Biotechnology
(Third Edition), 2019

2.59.3.1.2 Fumaric Acid


Fumaric acid is currently produced via isomerization of maleic acid, which is pro-
duced from maleic anhydride. The fermentative production of fumaric acid has a
high theoretical glucose to fumaric acid production yield (1.29 g g−1) when CO2
is provided. Fermentative production of fumaric acid has been investigated using
various fungal strains of Rhizopus sp. R. oryzae and R. arrhizus are the two most
widely studied fungi for fumaric acid production using mainly glucose and starch
hydrolysates. Koutinas et al.1 reported that fumaric acid concentrations up to 130 g-
 L−1 and productivities up to 4.25 g L−1 h have been achieved. Fumaric acid is currently
mainly used in the food industry as acidulant and the chemical industry for the
production of plasticisers and resins among other products.

> Read full chapter

Chain Polymerization of Vinyl


Monomers
R. Faust, in Polymer Science: A Comprehensive Reference, 2012

3.15.7.2.2(ii) -Pinene
The first example of living cationic isomerization polymerization of -pinene was
reported with the HCl–2-chloroethyl vinyl ether adduct [CH3CH(OCH2-CH2Cl)Cl] or
1-phenylethyl chloride/TiCl3(OiPr)-initiating system in the presence of nBu4NCl in
CH2Cl2 at –40 and –78 °C.76,77 The polymerization was rather slow even at relatively
high initiator (20 mM) and coinitiator (100 mM) concentrations. The much stronger
Lewis acid coinitiator TiCl4 induced an extremely rapid polymerization yielding
polymers with controlled molecular weight but with broad MWDs. The 1H NMR
analysis of the polymers showed a tert-chloride end group, and isomerized -pinene
repeat units with a cyclohexene ring. Copolymerization of -pinene with IB indicated
that the two monomers exhibit almost equal reactivity.78
> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like