Вы находитесь на странице: 1из 36

Catalysis

Solid heterogeneous catalysts such as in automobile catalytic converters are plated on structures
designed to maximize their surface area.

A low-temperature oxidation catalyst used to convert carbon monoxide to less toxic carbon
dioxide at room temperature. It can also remove formaldehyde from the air.

Catalysis is the increase in rate of a chemical reaction due to the participation of a substance
called a catalyst. Unlike other reagents in the chemical reaction, a catalyst is not consumed. A
catalyst may participate in multiple chemical transformations. The effect of a catalyst may vary
due to the presence of other substances known as inhibitors or poisons (which reduce the
catalytic activity) or promoters (which increase the activity).

Catalyzed reactions have a lower activation energy (rate-limiting free energy of activation) than
the corresponding uncatalyzed reaction, resulting in a higher reaction rate at the same
temperature. However, the mechanistic explanation of catalysis is complex. Catalysts may affect
the reaction environment favorably, or bind to the reagents to polarize bonds, e.g. acid catalysts
for reactions of carbonyl compounds, or form specific intermediates that are not produced
naturally, such as osmate esters in osmium tetroxide-catalyzed dihydroxylation of alkenes, or
cause dissociation of reagents to reactive forms, such as chemisorbed hydrogen in catalytic
hydrogenation.
Kinetically, catalytic reactions are typical chemical reactions; i.e. the reaction rate depends on
the frequency of contact of the reactants in the rate-determining step. Usually, the catalyst
participates in this slowest step, and rates are limited by amount of catalyst and its "activity". In
heterogeneous catalysis, the diffusion of reagents to the surface and diffusion of products from
the surface can be rate determining. A nanomaterial-based catalyst is an example of a
heterogeneous catalyst. Analogous events associated with substrate binding and product
dissociation apply to homogeneous catalysts.

Although catalysts are not consumed by the reaction itself, they may be inhibited, deactivated, or
destroyed by secondary processes. In heterogeneous catalysis, typical secondary processes
include coking where the catalyst becomes covered by polymeric side products. Additionally,
heterogeneous catalysts can dissolve into the solution in a solid–liquid system or evaporate in a
solid–gas system.

Background
The production of most industrially important chemicals involves catalysis. Similarly, most
biochemically significant processes are catalysed. Research into catalysis is a major field in
applied science and involves many areas of chemistry, notably in organometallic chemistry and
materials science. Catalysis is relevant to many aspects of environmental science, e.g. the
catalytic converter in automobiles and the dynamics of the ozone hole. Catalytic reactions are
preferred in environmentally friendly green chemistry due to the reduced amount of waste
generated,[1] as opposed to stoichiometric reactions in which all reactants are consumed and more
side products are formed. The most common catalyst is the hydrogen ion (H+). Many transition
metals and transition metal complexes are used in catalysis as well. Catalysts called enzymes are
important in biology.

A catalyst works by providing an alternative reaction pathway to the reaction product. The rate
of the reaction is increased as this alternative route has a lower activation energy than the
reaction route not mediated by the catalyst. The disproportionation of hydrogen peroxide creates
water and oxygen, as shown below.

2 H2O2 → 2 H2O + O2

This reaction is preferable in the sense that the reaction products are more stable than the starting
material, though the uncatalysed reaction is slow. In fact, the decomposition of hydrogen
peroxide is so slow that hydrogen peroxide solutions are commercially available. This reaction is
strongly affected by catalysts such as manganese dioxide, or the enzyme peroxidase in
organisms. Upon the addition of a small amount of manganese dioxide, the hydrogen peroxide
reacts rapidly. This effect is readily seen by the effervescence of oxygen.[2] The manganese
dioxide is not consumed in the reaction, and thus may be recovered unchanged, and re-used
indefinitely. Accordingly, manganese dioxide catalyses this reaction.[3]
Typical mechanism[

Catalysts generally react with one or more reactants to form intermediates that subsequently give
the final reaction product, in the process regenerating the catalyst. The following is a typical
reaction scheme, where C represents the catalyst, X and Y are reactants, and Z is the product of
the reaction of X and Y:

X + C → XC (1)
Y + XC → XYC (2)
XYC → CZ (3)
CZ → C + Z (4)

Although the catalyst is consumed by reaction 1, it is subsequently produced by reaction 4, so for


the overall reaction:

X+Y→Z

As a catalyst is regenerated in a reaction, often only small amounts are needed to increase the
rate of the reaction. In practice, however, catalysts are sometimes consumed in secondary
processes.

As an example of this process, in 2008 Danish researchers first revealed the sequence of events
when oxygen and hydrogen combine on the surface of titanium dioxide (TiO2, or titania) to
produce water. With a time-lapse series of scanning tunneling microscopy images, they
determined the molecules undergo adsorption, dissociation and diffusion before reacting. The
intermediate reaction states were: HO2, H2O2, then H3O2 and the final reaction product (water
molecule dimers), after which the water molecule desorbs from the catalyst surface.[4][5]

Reaction energetics[

Generic potential energy diagram showing the effect of a catalyst in a hypothetical exothermic
chemical reaction X + Y to give Z. The presence of the catalyst opens a different reaction
pathway (shown in red) with a lower activation energy. The final result and the overall
thermodynamics are the same.
Catalysts work by providing an (alternative) mechanism involving a different transition state and
lower activation energy. Consequently, more molecular collisions have the energy needed to
reach the transition state. Hence, catalysts can enable reactions that would otherwise be blocked
or slowed by a kinetic barrier. The catalyst may increase reaction rate or selectivity, or enable the
reaction at lower temperatures. This effect can be illustrated with a Boltzmann distribution and
energy profile diagram.

In the catalyzed elementary reaction, catalysts do not change the extent of a reaction: they have
no effect on the chemical equilibrium of a reaction because the rate of both the forward and the
reverse reaction are both affected (see also thermodynamics). The fact that a catalyst does not
change the equilibrium is a consequence of the second law of thermodynamics. Suppose there
was such a catalyst that shifted an equilibrium. Introducing the catalyst to the system would
result in reaction to move to the new equilibrium, producing energy. Production of energy is a
necessary result since reactions are spontaneous if and only if Gibbs free energy is produced, and
if there is no energy barrier, there is no need for a catalyst. Then, removing the catalyst would
also result in reaction, producing energy; i.e. the addition and its reverse process, removal, would
both produce energy. Thus, a catalyst that could change the equilibrium would be a perpetual
motion machine, a contradiction to the laws of thermodynamics.[6]

If a catalyst does change the equilibrium, then it must be consumed as the reaction proceeds, and
thus it is also a reactant. Illustrative is the base-catalysed hydrolysis of esters, where the
produced carboxylic acid immediately reacts with the base catalyst and thus the reaction
equilibrium is shifted towards hydrolysis.

The SI derived unit for measuring the catalytic activity of a catalyst is the katal, which is moles
per second. The productivity of a catalyst can be described by the turn over number (or TON)
and the catalytic activity by the turn over frequency (TOF), which is the TON per time unit. The
biochemical equivalent is the enzyme unit. For more information on the efficiency of enzymatic
catalysis, see the article on Enzymes.

The catalyst stabilizes the transition state more than it stabilizes the starting material. It decreases
the kinetic barrier by decreasing the difference in energy between starting material and transition
state. It does not change the energy difference between starting materials and products
(thermodynamic barrier), or the available energy (this is provided by the environment as heat or
light).

Materials

The chemical nature of catalysts is as diverse as catalysis itself, although some generalizations
can be made. Proton acids are probably the most widely used catalysts, especially for the many
reactions involving water, including hydrolysis and its reverse. Multifunctional solids often are
catalytically active, e.g. zeolites, alumina, higher-order oxides, graphitic carbon, nanoparticles,
nanodots, and facets of bulk materials. Transition metals are often used to catalyze redox
reactions (oxidation, hydrogenation). Examples are nickel, such as Raney nickel for
hydrogenation, and vanadium(V) oxide for oxidation of sulfur dioxide into sulfur trioxide. Many
catalytic processes, especially those used in organic synthesis, require so called "late transition
metals", which include palladium, platinum, gold, ruthenium, rhodium, and iridium.

Some so-called catalysts are really precatalysts. Precatalysts convert to catalysts in the reaction.
For example, Wilkinson's catalyst RhCl(PPh 3)3 loses one triphenylphosphine ligand before
entering the true catalytic cycle. Precatalysts are easier to store but are easily activated in situ.
Because of this preactivation step, many catalytic reactions involve an induction period.

Chemical species that improve catalytic activity are called co-catalysts (cocatalysts) or
promotors in cooperative catalysis.

Types
Catalysts can be heterogeneous or homogeneous, depending on whether a catalyst exists in the
same phase as the substrate. Biocatalysts (enzymes) are often seen as a separate group.

Heterogeneous catalysts

Main article: Heterogeneous catalysis

The microporous molecular structure of the zeolite ZSM-5 is exploited in catalysts used in
refineries
Zeolites are extruded as pellets for easy handling in catalytic reactors.

Heterogeneous catalysts act in a different phase than the reactants. Most heterogeneous catalysts
are solids that act on substrates in a liquid or gaseous reaction mixture. Diverse mechanisms for
reactions on surfaces are known, depending on how the adsorption takes place (Langmuir-
Hinshelwood, Eley-Rideal, and Mars-van Krevelen).[7] The total surface area of solid has an
important effect on the reaction rate. The smaller the catalyst particle size, the larger the surface
area for a given mass of particles.

For example, in the Haber process, finely divided iron serves as a catalyst for the synthesis of
ammonia from nitrogen and hydrogen. The reacting gases adsorb onto "active sites" on the iron
particles. Once adsorbed, the bonds within the reacting molecules are weakened, and new bonds
between the resulting fragments form in part due to their close proximity. In this way the
particularly strong triple bond in nitrogen is weakened and the hydrogen and nitrogen atoms
combine faster than would be the case in the gas phase, so the rate of reaction increases.[citation
needed]
Another place where a heterogeneous catalyst is applied is in the contact process (oxidation
of sulfur dioxide on vanadium(V) oxide for the production of sulfuric acid).

Heterogeneous catalysts are typically “supported,” which means that the catalyst is dispersed on
a second material that enhances the effectiveness or minimizes their cost. Sometimes the support
is merely a surface on which the catalyst is spread to increase the surface area. More often, the
support and the catalyst interact, affecting the catalytic reaction. Supports are porous materials
with a high surface area, most commonly alumina or various kinds of activated carbon.
Specialized supports include silicon dioxide, titanium dioxide, calcium carbonate, and barium
sulfate.

Electrocatalysts

In the context of electrochemistry, specifically in fuel cell engineering, various metal-containing


catalysts are used to enhance the rates of the half reactions that comprise the fuel cell. One
common type of fuel cell electrocatalyst is based upon nanoparticles of platinum that are
supported on slightly larger carbon particles. When in contact with one of the electrodes in a fuel
cell, this platinum increases the rate of oxygen reduction either to water, or to hydroxide or
hydrogen peroxide.

Homogeneous catalysts

Homogeneous catalysts function in the same phase as the reactants, but the mechanistic
principles invoked in heterogeneous catalysis are generally applicable. Typically homogeneous
catalysts are dissolved in a solvent with the substrates. One example of homogeneous catalysis
involves the influence of H+ on the esterification of carboxylic acids, such as the formation of
methyl acetate from acetic acid and methanol.[8] For inorganic chemists, homogeneous catalysis
is often synonymous with organometallic catalysts.[9]
: Organocatalysis

Whereas transition metals sometimes attract most of the attention in the study of catalysis, small
organic molecules without metals can also exhibit catalytic properties, as is apparent from the
fact that many enzymes lack transition metals. Typically, organic catalysts require a higher
loading (amount of catalyst per unit amount of reactant, expressed in mol% amount of substance)
than transition metal(-ion)-based catalysts, but these catalysts are usually commercially available
in bulk, helping to reduce costs. In the early 2000s, these organocatalysts were considered "new
generation" and are competitive to traditional metal(-ion)-containing catalysts. Organocatalysts
are supposed to operate akin to metal-free enzymes utilizing, e.g., non-covalent interactions such
as hydrogen bonding. The discipline organocatalysis is divided in the application of covalent
(e.g., proline, DMAP) and non-covalent (e.g., thiourea organocatalysis) organocatalysts referring
to the preferred catalyst-substrate binding and interaction, respectively.

Significance

Left: Partially caramelised cube sugar, Right: burning cube sugar with ash as catalyst

Estimates are that 90% of all commercially produced chemical products involve catalysts at
some stage in the process of their manufacture.[10] In 2005, catalytic processes generated about
$900 billion in products worldwide.[11] Catalysis is so pervasive that subareas are not readily
classified. Some areas of particular concentration are surveyed below.

Energy processing[

Petroleum refining makes intensive use of catalysis for alkylation, catalytic cracking (breaking
long-chain hydrocarbons into smaller pieces), naphtha reforming and steam reforming
(conversion of hydrocarbons into synthesis gas). Even the exhaust from the burning of fossil
fuels is treated via catalysis: Catalytic converters, typically composed of platinum and rhodium,
break down some of the more harmful byproducts of automobile exhaust.

2 CO + 2 NO → 2 CO2 + N2

With regard to synthetic fuels, an old but still important process is the Fischer-Tropsch synthesis
of hydrocarbons from synthesis gas, which itself is processed via water-gas shift reactions,
catalysed by iron. Biodiesel and related biofuels require processing via both inorganic and
biocatalysts.

Fuel cells rely on catalysts for both the anodic and cathodic reactions.
Catalytic heaters generate flameless heat from a supply of combustible fuel.

Bulk chemicals

Some of the largest-scale chemicals are produced via catalytic oxidation, often using oxygen.
Examples include nitric acid (from ammonia), sulfuric acid (from sulfur dioxide to sulfur
trioxide by the chamber process), terephthalic acid from p-xylene, and acrylonitrile from propane
and ammonia.

Many other chemical products are generated by large-scale reduction, often via hydrogenation.
The largest-scale example is ammonia, which is prepared via the Haber process from nitrogen.
Methanol is prepared from carbon monoxide.

Bulk polymers derived from ethylene and propylene are often prepared via Ziegler-Natta
catalysis. Polyesters, polyamides, and isocyanates are derived via acid-base catalysis.

Most carbonylation processes require metal catalysts, examples include the Monsanto acetic acid
process and hydroformylation.

Fine chemicals

Many fine chemicals are prepared via catalysis; methods include those of heavy industry as well
as more specialized processes that would be prohibitively expensive on a large scale. Examples
include olefin metathesis using Grubbs' catalyst, the Heck reaction, and Friedel-Crafts reactions.

Because most bioactive compounds are chiral, many pharmaceuticals are produced by
enantioselective catalysis (catalytic asymmetric synthesis).

Food processing[

One of the most obvious applications of catalysis is the hydrogenation (reaction with hydrogen
gas) of fats using nickel catalyst to produce margarine.[12] Many other foodstuffs are prepared via
biocatalysis (see below).

Biology

: Biocatalysis

In nature, enzymes are catalysts in metabolism and catabolism. Most biocatalysts are protein-
based, i.e. enzymes, but other classes of biomolecules also exhibit catalytic properties including
ribozymes, and synthetic deoxyribozymes.[13]

Biocatalysts can be thought of as intermediate between homogenous and heterogeneous


catalysts, although strictly speaking soluble enzymes are homogeneous catalysts and membrane-
bound enzymes are heterogeneous. Several factors affect the activity of enzymes (and other
catalysts) including temperature, pH, concentration of enzyme, substrate, and products. A
particularly important reagent in enzymatic reactions is water, which is the product of many
bond-forming reactions and a reactant in many bond-breaking processes.

Enzymes are employed to prepare many commodity chemicals including high-fructose corn
syrup and acrylamide.

Environment

Catalysis impacts the environment by increasing the efficiency of industrial processes, but
catalysis also plays a direct role in the environment. A notable example is the catalytic role of
chlorine free radicals in the breakdown of ozone. These radicals are formed by the action of
ultraviolet radiation on chlorofluorocarbons (CFCs).

Cl· + O3 → ClO· + O2
ClO· + O· → Cl· + O2

History
In a general sense,[14] anything that increases the rate of a process is a "catalyst", a term derived
from Greek καταλύειν, meaning "to annul," or "to untie," or "to pick up." The term catalysis was
coined by Jöns Jakob Berzelius in 1835[15] to describe reactions that are accelerated by
substances that remain unchanged after the reaction. Other early chemists involved in catalysis
were Eilhard Mitscherlich[16] who referred to contact processes and Johann Wolfgang
Döbereiner[17] who spoke of contact action and whose lighter based on hydrogen and a platinum
sponge became a huge commercial success in the 1820s. Humphry Davy discovered the use of
platinum in catalysis.[18] In the 1880s, Wilhelm Ostwald at Leipzig University started a
systematic investigation into reactions that were catalyzed by the presence of acids and bases,
and found that chemical reactions occur at finite rates and that these rates can be used to
determine the strengths of acids and bases. For this work, Ostwald was awarded the 1909 Nobel
Prize in Chemistry.[19]

Inhibitors, poisons and promoters


Substances that reduce the action of catalysts are called catalyst inhibitors if reversible, and
catalyst poisons if irreversible. Promoters are substances that increase the catalytic activity, even
though they are not catalysts by themselves.

Inhibitors are sometimes referred to as "negative catalysts" since they decrease the reaction
rate.[20] However the term inhibitor is preferred since they do not work by introducing a reaction
path with higher activation energy; this would not reduce the rate since the reaction would
continue to occur by the non-catalyzed path. Instead they act either by deactivating catalysts, or
by removing reaction intermediates such as free radicals.[20][21]

The inhibitor may modify selectivity in addition to rate. For instance, in the reduction of ethyne
to ethene, the catalyst is palladium (Pd) partly "poisoned" with lead(II) acetate (Pb(CH3COO)2).
Without the deactivation of the catalyst, the ethene produced will be further reduced to
ethane.[22][23]

The inhibitor can produce this effect by e.g. selectively poisoning only certain types of active
sites. Another mechanism is the modification of surface geometry. For instance, in
hydrogenation operations, large planes of metal surface function as sites of hydrogenolysis
catalysis while sites catalyzing hydrogenation of unsaturates are smaller. Thus, a poison that
covers surface randomly will tend to reduce the number of uncontaminated large planes but leave
proportionally more smaller sites free, thus changing the hydrogenation vs. hydrogenolysis
selectivity. Many other mechanisms are also possible.

Promoters can cover up surface to prevent production of a mat of coke, or even actively remove
such material (e.g. rhenium on platinum in platforming). They can aid the dispersion of the
catalytic material or bind to reagents.

Current market
The global demand on catalysts in 2010 was estimated at approximately 29.5 billions USD. With
the rapid recovery in automotive and chemical industry overall, the global catalyst market is
expected to experience fast growth in the next years.[24]

References[IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book")


(1997). Online corrected version: (2006–) "catalyst".

1. ^ "The 12 Principles of Green Chemistry". United States Environmental Protection Agency. Retrieved
2012-04-30.
2. ^ "Genie in a Bottle". University of Minnesota. 2005-03-02.
3. ^ Richard I. Masel “Chemical Kinetics and Catalysis” Wiley-Interscience, New York, 2001. ISBN 0-471-
24197-0.
4. ^ Jacoby, Mitch (16 February 2009). "Making Water Step by Step". Chemical & Engineering News. p. 10.
5. ^ Matthiesen J, Wendt S, Hansen JØ, Madsen GK, Lira E, Galliker P, Vestergaard EK, Schaub R,
Laegsgaard E, Hammer B, Besenbacher F (2009). "Observation of All the Intermediate Steps of a Chemical
Reaction on an Oxide Surface by Scanning Tunneling Microscopy". ACS Nano 3 (3): 517–526.
doi:10.1021/nn8008245. ISSN 1520-605X. PMID 19309169.
6. ^ A.J.B. Robertson Catalysis of Gas Reactions by Metals. Logos Press, London, 1970.
7. ^ Helmut Knözinger, Karl Kochloefl “Heterogeneous Catalysis and Solid Catalysts” in Ullmann's
Encyclopedia of Industrial Chemistry 2002, Wiley-VCH, Weinheim. doi:10.1002/14356007.a05_313.
Article Online Posting Date: January 15, 2003
8. ^ Arno Behr “Organometallic Compounds and Homogeneous Catalysis” Ullmann's Encyclopedia of
Industrial Chemistry, 2002, Wiley-VCH, Weinheim. doi:10.1002/14356007.a18_215. Article Online
Posting Date: June 15, 2000
9. ^ C. Elschenbroich, Organometallics (2006) Wiley-VCH: Weinheim. ISBN 978-3-527-29390-2
10. ^ "Recognizing the Best in Innovation: Breakthrough Catalyst". R&D Magazine, September 2005, p. 20.
11. ^ 1.4.3 INDUSTRIAL PROCESS EFFICIENCY. climatetechnology.gov
12. ^ "Types of catalysis". Chemguide. Retrieved 2008-07-09.
13. ^ D. L. Nelson and M. M. Cox Lehninger, Principles of Biochemistry 3rd Ed. Worth Publishing: New
York, 2000. ISBN 1-57259-153-6.
14. ^ Bård Lindström and Lars J. Petterson (2003) "A brief history of catalysis," Cattech, 7 (4) : 130-138.
Available on-line at: ScienceNet.
15. ^ J. J. Berzelius, Årsberättelsen om framsteg i fysik och kemi [Annual report on progress in physics and
chemistry], (Stockholm, Sweden: Royal Swedish Academy of Sciences, 1835). After reviewing Eilhard
Mitscherlich's research on the formation of ether, Berzelius coins the word katalys (catalysis) on page 245:

Original: Jag skall derföre, för att begagna en i kemien välkänd härledning, kalla den kroppars katalytiska
kraft, sönderdelning genom denna kraft katalys, likasom vi med ordet analys beteckna åtskiljandet af
kroppars beståndsdelar medelst den vanliga kemiska frändskapen.

Translation: I shall, therefore, to employ a well-known derivation in chemistry, call [the catalytic]
bodies [i.e., substances] the catalytic force and the decomposition of [other] bodies by this force
catalysis, just as we signify by the word analysis the separation of the constituents of bodies by
the usual chemical affinities.

16. ^ E. Mitscherlich (1834) "Ueber die Aetherbildung" (On the formation of ether), Annalen der Physik und
Chemie, 31 (18) : 273-282.
17. ^ See:
 Döbereiner (1822) "Glühendes Verbrennen des Alkohols durch verschiedene erhitzte Metalle und
Metalloxyde" (Incandescent burning of alcohol by various heated metals and metal oxides),
Journal für Chemie und Physik, 34 : 91-92.
 Döbereiner (1823) "Neu entdeckte merkwürdige Eigenschaften des Platinsuboxyds, des oxydirten
Schwefel-Platins und des metallischen Platinstaubes" (Newly discovered remarkable properties of
platinum suboxide, oxidized platinum sulfide and metallic platinum dust), Journal für Chemie und
Physik, 38 : 321-326.
18. ^ Sir Humphry Davy (1817) "Some new experiments and observations on the combustion of gaseous
mixtures, with an account of a method of preserving a continued light in mixtures of inflammable gases and
air without flame," Philosophical Transactions of the Royal Society of London, 107 : 77-85.
19. ^ M.W. Roberts (2000). "Birth of the catalytic concept (1800–1900)". Catalysis Letters 67 (1): 1–4.
doi:10.1023/A:1016622806065.
20. ^ a b K.J. Laidler Physical Chemistry with Biological Applications, Benjamin/Cummings 1978, p.415-417
21. ^ K.J. Laidler and J.H. Meiser, Physical Chemistry, Benjamin/Cummings (1982), p.425
22. ^ W.P. Jencks, “Catalysis in Chemistry and Enzymology” McGraw-Hill, New York, 1969. ISBN 0-07-
032305-4
23. ^ Myron L Bender, Makoto Komiyama, Raymond J Bergeron “The Bioorganic Chemistry of Enzymatic
Catalysis” Wiley-Interscience, Hoboken, U.S., 1984 ISBN 0-471-05991-9
24. ^ "Market Report: Global Catalyst Market, 2nd Edition". Acmite Market Intelligence.

 Science Aid: Catalysts Page for high school level science


 W.A. Herrmann Technische Universität presentation
 Alumite Catalyst, Kameyama-Sakurai Laboratory, Japan
 Inorganic Chemistry and Catalysis Group, Utrecht University, The Netherlands
 Centre for Surface Chemistry and Catalysis
 Carbons & Catalysts Group, University of Concepcion, Chile
 Center for Enabling New Technologies Through Catalysis, An NSF Center for Chemical
Innovation, USA
 "Bubbles turn on chemical catalysts", Science News magazine online, April 6, 2009.
Acid catalysis

In acid-catalyzed Fischer esterification, the proton binds to oxygens and functions as a Lewis
acid to activate the ester carbonyl (top row) as an electrophile, and converts the hydroxyl into the
good leaving group water (bottom left). Both lower the kinetic barrier and speed up the
attainment of chemical equilibrium.In acid catalysis and base catalysis a chemical reaction is
catalyzed by an acid or a base. The acid is often the proton and the base is often a hydroxide ion.
Typical reactions catalyzed by proton transfer are esterfications and aldol reactions. In these
reactions the conjugate acid of the carbonyl group is a better electrophile than the neutral
carbonyl group itself. Catalysis by either acid or base can occur in two different ways: specific
catalysis and general catalysis.

Use in synthesis
Acid catalysis is mainly used for organic chemical reactions. There are many possible chemical
compounds that can act as sources for the protons to be transferred in an acid catalysis system. A
compound such as sulfuric acid, H2SO4, can be used. Usually this is done to create a more likely
leaving group, such as converting an OH group to a H2O+ group, which can then be eliminated as
water (H2O). Acids specifically used for acid catalysis include hydrofluoric acid (in the
alkylation process), phosphoric acid, toluenesulfonic acid, polystyrene sulfonate, heteropoly
acids, zeolites and graphene oxide.

With carbonyl compounds such as esters, synthesis and hydrolysis go through a tetrahedral
transition state, where the central carbon has an oxygen, an alcohol group, and the original alkyl
group. Strong acids protonate the carbonyl, which makes the oxygen positively charged, so that
it can easily receive the double bond electrons when the alcohol attacks the carbonyl carbon.
This enables ester synthesis and hydrolysis. The reaction is an equilibrium between the ester and
its cleavage to carboxylic acid and alcohol. On the contrary, strong bases deprotonate the
attacking alcohol or amine, which also promotes the reaction. However, bases also deprotonate
the acid, which is irreversible. Therefore, in a strongly basic, aqueous environment, esters only
hydrolyze.

Kinetics
Specific catalysis

In specific acid catalysis taking place in solvent S, the reaction rate is proportional to the
concentration of the protonated solvent molecules SH+.[1] The acid catalyst itself (AH) only
contributes to the rate acceleration by shifting the chemical equilibrium between solvent S and
AH in favor of the SH+ species.

S + AH → SH+ + A-

For example in an aqueous buffer solution the reaction rate for reactants R depends on the pH of
the system but not on the concentrations of different acids.

This type of chemical kinetics is observed when reactant R1 is in a fast equilibrium with its
conjugate acid R1H+ which proceeds to react slowly with R2 to the reaction product; for
example, in the acid catalysed aldol reaction.

General catalysis

In general acid catalysis all species capable of donating protons contribute to reaction rate
acceleration.[2] The strongest acids are most effective. Reactions in which proton transfer is rate-
determining exhibit general acid catalysis, for example diazonium coupling reactions.

When keeping the pH at a constant level but changing the buffer concentration a change in rate
signals a general acid catalysis. A constant rate is evidence for a specific acid catalyst.

References^ "IUPAC Compendium of Chemical Terminology, 2nd Edition (1997)".


www.iupac.org. Retrieved 2009-11-22.

1. ^ "IUPAC Compendium of Chemical Terminology, 2nd Edition (1997)". www.iupac.org.


Retrieved 2009-11-22.
Adsorption
.

Adsorption is the adhesion of atoms, ions, or molecules from a gas, liquid, or dissolved solid to
a surface.[1] This process creates a film of the adsorbate on the surface of the adsorbent. This
process differs from absorption, in which a fluid (the absorbate) permeates or is dissolved by a
liquid or solid (the absorbent).[2] Note that adsorption is a surface-based process while absorption
involves the whole volume of the material. The term sorption encompasses both processes, while
desorption is the reverse of adsorption. It is a surface phenomenon.

IUPAC definition

Increase in the concentration of a substance at the interface


of a condensed and a liquid or gaseous layer owing to the operation
of surface forces.

Note 1: Adsorption of proteins is of great importance when a material


is in contact with blood or body fluids. In the case of blood, albumin,
which is largely predominant, is generally adsorbed first, and then
rearrangements occur in favor of other minor proteins according to
surface affinity against mass law selection (Vroman effect).

Note 2: Adsorbed molecules are those that are resistant to washing


with the same solvent medium in the case of adsorption from
solutions. The washing conditions can thus modify the measurement
results, particularly when the interaction energy is low. [3]

Similar to surface tension, adsorption is a consequence of surface energy. In a bulk material, all
the bonding requirements (be they ionic, covalent, or metallic) of the constituent atoms of the
material are filled by other atoms in the material. However, atoms on the surface of the adsorbent
are not wholly surrounded by other adsorbent atoms and therefore can attract adsorbates. The
exact nature of the bonding depends on the details of the species involved, but the adsorption
process is generally classified as physisorption (characteristic of weak van der Waals forces) or
chemisorption (characteristic of covalent bonding). It may also occur due to electrostatic
attraction.[4]

Adsorption is present in many natural physical, biological, and chemical systems, and is widely
used in industrial applications such as activated charcoal, capturing and using waste heat to
provide cold water for air conditioning and other process requirements (adsorption chillers),
synthetic resins, increase storage capacity of carbide-derived carbons, and water purification.
Adsorption, ion exchange, and chromatography are sorption processes in which certain
adsorbates are selectively transferred from the fluid phase to the surface of insoluble, rigid
particles suspended in a vessel or packed in a column. Lesser known, are the pharmaceutical
industry applications as a means to prolong neurological exposure to specific drugs or parts
thereof.

The word "adsorption" was coined in 1881 by German physicist Heinrich Kayser (1853-1940).[5]

Isotherms[
Adsorption is usually described through isotherms, that is, the amount of adsorbate on the
adsorbent as a function of its pressure (if gas) or concentration (if liquid) at constant temperature.
The quantity adsorbed is nearly always normalized by the mass of the adsorbent to allow
comparison of different materials.

: Freundlich equation
The first mathematical fit to an isotherm was published by Freundlich and Küster (1894) and is a
purely empirical formula for gaseous adsorbates,

where is the quantity adsorbed, is the mass of the adsorbent, is the pressure of adsorbate
and and are empirical constants for each adsorbent-adsorbate pair at a given temperature. The
function is not adequate at very high pressure because in reality has an asymptotic
maximum as pressure increases without bound. As the temperature increases, the constants and
change to reflect the empirical observation that the quantity adsorbed rises more slowly and
higher pressures are required to saturate the surface.

Langmuir[

Langmuir equation

Irving Langmuir was the first to derive a scientifically based adsorption isotherm in 1918.[6] The
model applies to gases adsorbed on solid surfaces. It is a semi-empirical isotherm with a kinetic
basis and was derived based on statistical thermodynamics. It is the most common isotherm
equation to use due to its simplicity and its ability to fit a variety of adsorption data. It is based
on four assumptions:

1. All of the adsorption sites are equivalent and each site can only accommodate one
molecule.
2. The surface is energetically homogeneous and adsorbed molecules do not interact.
3. There are no phase transitions.
4. At the maximum adsorption, only a monolayer is formed. Adsorption only occurs on
localized sites on the surface, not with other adsorbates.
These four assumptions are seldom all true: there are always imperfections on the surface,
adsorbed molecules are not necessarily inert, and the mechanism is clearly not the same for the
very first molecules to adsorb to a surface as for the last. The fourth condition is the most
troublesome, as frequently more molecules will adsorb to the monolayer; this problem is
addressed by the BET isotherm for relatively flat (non-microporous) surfaces. The Langmuir
isotherm is nonetheless the first choice for most models of adsorption, and has many applications
in surface kinetics (usually called Langmuir–Hinshelwood kinetics) and thermodynamics.

Langmuir suggested that adsorption takes place through this mechanism: ,


where A is a gas molecule and S is an adsorption site. The direct and inverse rate constants are k
and k−1. If we define surface coverage, , as the fraction of the adsorption sites occupied, in the
equilibrium we have:

or

where is the partial pressure of the gas or the molar concentration of the solution. For very
low pressures and for high pressures

is difficult to measure experimentally; usually, the adsorbate is a gas and the quantity adsorbed
is given in moles, grams, or gas volumes at standard temperature and pressure (STP) per gram of
adsorbent. If we call vmon the STP volume of adsorbate required to form a monolayer on the

adsorbent (per gram of adsorbent), and we obtain an expression for a straight line:

Through its slope and y-intercept we can obtain vmon and K, which are constants for each
adsorbent/adsorbate pair at a given temperature. vmon is related to the number of adsorption sites
through the ideal gas law. If we assume that the number of sites is just the whole area of the solid
divided into the cross section of the adsorbate molecules, we can easily calculate the surface area
of the adsorbent. The surface area of an adsorbent depends on its structure; the more pores it has,
the greater the area, which has a big influence on reactions on surfaces.

If more than one gas adsorbs on the surface, we define as the fraction of empty sites and we
have:
Also, we can define as the fraction of the sites occupied by the j-th gas:

where i is each one of the gases that adsorb.

BET

BET theory

Often molecules do form multilayers, that is, some are adsorbed on already adsorbed molecules
and the Langmuir isotherm is not valid. In 1938 Stephen Brunauer, Paul Emmett, and Edward
Teller developed a model isotherm that takes that possibility into account. Their theory is called
BET theory, after the initials in their last names. They modified Langmuir's mechanism as
follows:

A(g) + S ⇌ AS
A(g) + AS ⇌ A2S
A(g) + A2S ⇌ A3S and so on

Langmuir isotherm (red) and BET isotherm (green)

The derivation of the formula is more complicated than Langmuir's (see links for complete
derivation). We obtain:
x is the pressure divided by the vapor pressure for the adsorbate at that temperature (usually
denoted ), v is the STP volume of adsorbed adsorbate, vmon is the STP volume of the
amount of adsorbate required to form a monolayer and c is the equilibrium constant K we used in
Langmuir isotherm multiplied by the vapor pressure of the adsorbate. The key assumption used
in deriving the BET equation that the successive heats of adsorption for all layers except the first
are equal to the heat of condensation of the adsorbate.

The Langmuir isotherm is usually better for chemisorption and the BET isotherm works better
for physisorption for non-microporous surfaces.

Kisliuk

Two adsorbate nitrogen molecules adsorbing onto a tungsten adsorbent from the precursor state
around an island of previously adsorbed adsorbate (left) and via random adsorption (right)

In other instances, molecular interactions between gas molecules previously adsorbed on a solid
surface form significant interactions with gas molecules in the gaseous phases. Hence, adsorption
of gas molecules to the surface is more likely to occur around gas molecules that are already
present on the solid surface, rendering the Langmuir adsorption isotherm ineffective for the
purposes of modelling. This effect was studied in a system where nitrogen was the adsorbate and
tungsten was the adsorbent by Paul Kisliuk (1922–2008) in 1957.[7] To compensate for the
increased probability of adsorption occurring around molecules present on the substrate surface,
Kisliuk developed the precursor state theory, whereby molecules would enter a precursor state at
the interface between the solid adsorbent and adsorbate in the gaseous phase. From here,
adsorbate molecules would either adsorb to the adsorbent or desorb into the gaseous phase. The
probability of adsorption occurring from the precursor state is dependent on the adsorbate’s
proximity to other adsorbate molecules that have already been adsorbed. If the adsorbate
molecule in the precursor state is in close proximity to an adsorbate molecule that has already
formed on the surface, it has a sticking probability reflected by the size of the SE constant and
will either be adsorbed from the precursor state at a rate of kEC or will desorb into the gaseous
phase at a rate of kES. If an adsorbate molecule enters the precursor state at a location that is
remote from any other previously adsorbed adsorbate molecules, the sticking probability is
reflected by the size of the SD constant.

These factors were included as part of a single constant termed a "sticking coefficient," kE,
described below:
As SD is dictated by factors that are taken into account by the Langmuir model, SD can be
assumed to be the adsorption rate constant. However, the rate constant for the Kisliuk model (R’)
is different to that of the Langmuir model, as R’ is used to represent the impact of diffusion on
monolayer formation and is proportional to the square root of the system’s diffusion coefficient.
The Kisliuk adsorption isotherm is written as follows, where Θ(t) is fractional coverage of the
adsorbent with adsorbate, and t is immersion time:

Solving for Θ(t) yields:

Adsorption enthalpy

Adsorption constants are equilibrium constants, therefore they obey van 't Hoff's equation:

As can be seen in the formula, the variation of K must be isosteric, that is, at constant coverage.
If we start from the BET isotherm and assume that the entropy change is the same for
liquefaction and adsorption we obtain

that is to say, adsorption is more exothermic than liquefaction.

Adsorbents
Characteristics and general requirements
Activated carbon is used as an adsorbent

Adsorbents are used usually in the form of spherical pellets, rods, moldings, or monoliths with
hydrodynamic diameters between 0.5 and 10 mm. They must have high abrasion resistance, high
thermal stability and small pore diameters, which results in higher exposed surface area and
hence high surface capacity for adsorption. The adsorbents must also have a distinct pore
structure that enables fast transport of the gaseous vapors.

Most industrial adsorbents fall into one of three classes:

 Oxygen-containing compounds – Are typically hydrophilic and polar, including materials


such as silica gel and zeolites.
 Carbon-based compounds – Are typically hydrophobic and non-polar, including materials
such as activated carbon and graphite.
 Polymer-based compounds – Are polar or non-polar functional groups in a porous
polymer matrix.

Silica gel

Silica gel is a chemically inert, nontoxic, polar and dimensionally stable (< 400 °C or 750 °F)
amorphous form of SiO2. It is prepared by the reaction between sodium silicate and acetic acid,
which is followed by a series of after-treatment processes such as aging, pickling, etc. These
after treatment methods results in various pore size distributions.

Silica is used for drying of process air (e.g. oxygen, natural gas) and adsorption of heavy (polar)
hydrocarbons from natural gas.

Zeolites

Zeolites are natural or synthetic crystalline aluminosilicates, which have a repeating pore
network and release water at high temperature. Zeolites are polar in nature.

They are manufactured by hydrothermal synthesis of sodium aluminosilicate or another silica


source in an autoclave followed by ion exchange with certain cations (Na+, Li+, Ca2+, K+, NH4+).
The channel diameter of zeolite cages usually ranges from 2 to 9 Å (200 to 900 pm). The ion
exchange process is followed by drying of the crystals, which can be pelletized with a binder to
form macroporous pellets.

Zeolites are applied in drying of process air, CO2 removal from natural gas, CO removal from
reforming gas, air separation, catalytic cracking, and catalytic synthesis and reforming.

Non-polar (siliceous) zeolites are synthesized from aluminum-free silica sources or by


dealumination of aluminum-containing zeolites. The dealumination process is done by treating
the zeolite with steam at elevated temperatures, typically greater than 500 °C (930 °F). This high
temperature heat treatment breaks the aluminum-oxygen bonds and the aluminum atom is
expelled from the zeolite framework.
Activated carbon

Activated carbon is a highly porous, amorphous solid consisting of microcrystallites with a


graphite lattice, usually prepared in small pellets or a powder. It is non-polar and cheap. One of
its main drawbacks is that it reacts with oxygen at moderate temperatures (over 300 °C).

Activated carbon nitrogen isotherm showing a marked microporous type I behavior

Activated carbon can be manufactured from carbonaceous material, including coal (bituminous,
subbituminous, and lignite), peat, wood, or nutshells (e.g., coconut). The manufacturing process
consists of two phases, carbonization and activation. The carbonization process includes drying
and then heating to separate by-products, including tars and other hydrocarbons from the raw
material, as well as to drive off any gases generated. The process is completed by heating the
material over 400 °C (750 °F) in an oxygen-free atmosphere that cannot support combustion.
The carbonized particles are then "activated" by exposing them to an oxidizing agent, usually
steam or carbon dioxide at high temperature. This agent burns off the pore blocking structures
created during the carbonization phase and so, they develop a porous, three-dimensional graphite
lattice structure. The size of the pores developed during activation is a function of the time that
they spend in this stage. Longer exposure times result in larger pore sizes. The most popular
aqueous phase carbons are bituminous based because of their hardness, abrasion resistance, pore
size distribution, and low cost, but their effectiveness needs to be tested in each application to
determine the optimal product.

Activated carbon is used for adsorption of organic substances and non-polar adsorbates and it is
also usually used for waste gas (and waste water) treatment. It is the most widely used adsorbent
since most of its chemical (e.g. surface groups) and physical properties (e.g. pore size
distribution and surface area) can be tuned according to what is needed. Its usefulness also
derives from its large micropore (and sometimes mesopore) volume and the resulting high
surface area.

Protein adsorption of biomaterials


Protein adsorption is a process that has a fundamental role in the field of biomaterials. Indeed,
biomaterial surfaces in contact with biological media, such as blood or serum, are immediately
coated by proteins. Therefore, living cells do not interact directly with the biomaterial surface,
but with the adsorbed proteins layer. This protein layer mediates the interaction between
biomaterials and cells, translating biomaterial physical and chemical properties into a "biological
language".[8] In fact, cell membrane receptors bind to protein layer bioactive sites and these
receptor-protein binding events are transduced, through the cell membrane, in a manner that
stimulates specific intracellular processes that then determine cell adhesion, shape, growth and
differentiation. Protein adsorption is influenced by many surface properties such as surface
wettability, surface chemical composition [9] and surface nanometre-scale morphology.[10]

Adsorption chillers
Combining an adsorbent with a refrigerant, adsorption chillers use heat to provide a cooling
effect. This heat, in the form of hot water, may come from any number of industrial sources
including waste heat from industrial processes, prime heat from solar thermal installations or
from the exhaust or water jacket heat of a piston engine or turbine.

Although there are similarities between absorption and adsorption refrigeration, the latter is
based on the interaction between gases and solids. The adsorption chamber of the chiller is filled
with a solid material (for example zeolite, silica gel, alumina, active carbon and certain types of
metal salts), which in its neutral state has adsorbed the refrigerant. When heated, the solid
desorbs (releases) refrigerant vapour, which subsequently is cooled and liquefied. This liquid
refrigerant then provides its cooling effect at the evaporator, by absorbing external heat and
turning back into a vapour. In the final stage the refrigerant vapour is (re)adsorbed into the
solid.[11] As an adsorption chiller requires no moving parts, it is relatively quiet.

Portal site mediated adsorption


Portal site mediated adsorption is a model for site-selective activated gas adsorption in metallic
catalytic systems that contain a variety of different adsorption sites. In such systems, low-
coordination "edge and corner" defect-like sites can exhibit significantly lower adsorption
enthalpies than high-coordination (basal plane) sites. As a result, these sites can serve as
"portals" for very rapid adsorption to the rest of the surface. The phenomenon relies on the
common "spillover" effect (described below), where certain adsorbed species exhibit high
mobility on some surfaces. The model explains seemingly inconsistent observations of gas
adsorption thermodynamics and kinetics in catalytic systems where surfaces can exist in a range
of coordination structures, and it has been successfully applied to bimetallic catalytic systems
where synergistic activity is observed.

In contrast to pure spillover, portal site adsorption refers to surface diffusion to adjacent
adsorption sites, not to non-adsorptive support surfaces.

The model appears to have been first proposed for carbon monoxide on silica-supported
platinum by Brandt et al. (1993).[12] A similar, but independent model was developed by King
and co-workers[13][14][15] to describe hydrogen adsorption on silica-supported alkali promoted
ruthenium, silver-ruthenium and copper-ruthenium bimetallic catalysts. The same group applied
the model to CO hydrogenation (Fischer–Tropsch synthesis).[16] Zupanc et al. (2002)
subsequently confirmed the same model for hydrogen adsorption on magnesia-supported
caesium-ruthenium bimetallic catalysts.[17] Trens et al. (2009) have similarly described CO
surface diffusion on carbon-supported Pt particles of varying morphology.[18]

Adsorption spillover
In the case catalytic or adsorbent systems where a metal species is dispersed upon a support (or
carrier) material (often quasi-inert oxides, such as alumina or silica), it is possible for an
adsorptive species to indirectly adsorb to the support surface under conditions where such
adsorption is thermodynamically unfavorable. The presence of the metal serves as a lower-
energy pathway for gaseous species to first adsorb to the metal and then diffuse on the support
surface. This is possible because the adsorbed species attains a lower energy state once it has
adsorbed to the metal, thus lowering the activation barrier between the gas phase species and the
support-adsorbed species.

Hydrogen spillover is the most common example of an adsorptive spillover. In the case of
hydrogen, adsorption is most often accompanied with dissociation of molecular hydrogen (H2) to
atomic hydrogen (H), followed by spillover of the hydrogen atoms present.

The spillover effect has been used to explain many observations in heterogeneous catalysis and
adsorption.[19]

Polymer adsorption
Adsorption of molecules onto polymer surfaces is central to a number of applications, including
development of non-stick coatings and in various biomedical devices. Polymers may also be
adsorbed to surfaces through polyelectrolyte adsorption.

Adsorption in viruses
Adsorption is the first step in the viral life cycle. The next steps are penetration, uncoating,
synthesis (transcription if needed, and translation), and release. The virus replication cycle, in
this respect, is similar for all types of viruses. Factors such as transcription may or may not be
needed if the virus is able to integrate its genomic information in the cell's nucleus, or if the virus
can replicate itself directly within the cell's cytoplasm.

In popular culture
The game of Tetris is a puzzle game in which blocks of 4 are adsorbed onto a surface during
game play. Scientists have used Tetris blocks "as a proxy for molecules with a complex shape"
and their "adsorption on a flat surface" for studying the thermodynamics of nanoparticles.[20][21]
References
"Glossary". The Brownfields and Land Revitalization Technology Support Center. Retrieved 2009-12-21.

1. "absorption (chemistry)". Memidex (WordNet) Dictionary/Thesaurus. Retrieved 2010-11-02.


2. Jump up "Glossary of atmospheric chemistry terms (Recommendations 1990)". Pure and Applied
Chemistry 62: 2167. 1990. doi:10.1351/goldbook.A00155.
3. Jump up ^ Ferrari, L.; Kaufmann, J.; Winnefeld, F.; Plank, J. (2010). "Interaction of cement model
systems with superplasticizers investigated by atomic force microscopy, zeta potential, and adsorption
measurements". J Colloid Interface Sci. 347 (1): 15–24. doi:10.1016/j.jcis.2010.03.005. PMID 20356605.
4. Jump up ^ Heinrich Kayser (1881) "Ueber die Verdichtung von Gasen an Oberflächen in ihrer
Abhängigkeit von Druck und Temperatur" (On the condensation of gases on surfaces in their dependence
on pressure and temperature), Annalen der Physik und Chemie, 3rd series, vol. 12 or 248 (4) : 526–537. In
this study of the adsorption of gases by charcoal, the first use of the word "adsorption" appears on page
527: "Schon Saussure kannte die beiden für die Grösse der Adsorption massgebenden Factoren, den Druck
und die Temperatur, da er Erniedrigung des Druckes oder Erhöhung der Temperatur zur Befreiung der
porösen Körper von Gasen benutzte." (Saussaure already knew the two factors that determine the quantity
of adsorption – [namely,] the pressure and temperature – since he used the lowering of the pressure or the
raising of the temperature to free the porous substances of gases.)
5. Jump up ^ Czepirski, L.; Balys, M. R.; Komorowska-Czepirska, E. (2000). "Some generalization of
Langmuir adsorption isotherm.". Internet Journal of Chemistry 3 (14).
6. Jump up ^ Kisliuk, P. (1957). "The sticking probabilities of gases chemisorbed on the surfaces of solids".
Journal of Physics and Chemistry of Solids 3 (1–2): 95–101. doi:10.1016/0022-3697(57)90054-9.
7. Jump up ^ Wilson, CJ; Clegg, RE; Leavesley, DI; Pearcy, MJ (2005). "Mediation of Biomaterial-Cell
Interactions by Adsorbed Proteins: A Review". Tissue engineering 11 (1): 1–18. doi:10.1089/ten.2005.11.1.
PMID 15738657.
8. Jump up ^ Sivaraman B., Fears K.P., Latour R.A. (2009). "Investigation of the effects of surface
chemistry and solution concentration on the conformation of adsorbed proteins using an improved circular
dichroism method". Langmuir 25 (5): 3050–6. doi:10.1021/la8036814. PMC 2891683. PMID 19437712.
9. Jump up ^ Scopelliti, Pasquale Emanuele; Borgonovo, Antonio; Indrieri, Marco; Giorgetti, Luca;
Bongiorno, Gero; Carbone, Roberta; Podestà, Alessandro; Milani, Paolo (2010). "The effect of surface
nanometre-scale morphology on protein adsorption". In Zhang, Shuguang. PLoS ONE 5 (7): e11862.
doi:10.1371/journal.pone.0011862.
10. Jump up ^ Pilatowsky, I.; Romero, R.J.; Isaza, C.A.; Gamboa, S.A.; Sebastian, P.J. and Rivera, W. (2011).
"Chapter 5: Sorption Refrigeration Systems". Cogeneration Fuel Cell-Sorption Air Conditioning Systems.
Green Energy and Technology. Springer. pp. 99,100. ISBN 978-1-84996-027-4. Retrieved 10 May 2011.
11. Jump up ^ Brandt, R. K.; Hughes, M. R.; Bourget, L. P.; Truszkowska, K.; Greenler, R. G. (1993). "The
interpretation of CO adsorbed on Pt/SiO2 of two different particle-size distributions". Surface Science 286
(1–2): 15–25. doi:10.1016/0039-6028(93)90552-U.
12. Jump up ^ Uner, D. O.; Savargoankar, N.; Pruski, M.; King, T. S. (1997). "The effects of alkali promoters
on the dynamics of hydrogen chemisorption and syngas reaction kinetics on Ru/SiO 2 catalysts". Studies in
Surface Science and Catalysis. Studies in Surface Science and Catalysis 109: 315–324. doi:10.1016/S0167-
2991(97)80418-1. ISBN 9780444826091.
13. Jump up ^ Narayan, R. L.; King, T. S. (1998). "Hydrogen adsorption states on silica-supported Ru-Ag and
Ru-Cu bimetallic catalysts investigated via microcalorimetry". Thermochimica Acta 312 (1–2): 105–114.
doi:10.1016/S0040-6031(97)00444-9.
14. Jump up ^ VanderWiel, D. P.; Pruski, M.; King, T. S. (1999). "A Kinetic Study of the Adsorption and
Reaction of Hydrogen on Silica-Supported Ruthenium and Silver-Ruthenium Bimetallic Catalysts during
the Hydrogenation of Carbon Monoxide". Journal of Catalysis 188 (1): 186–202.
doi:10.1006/jcat.1999.2646.
15. Jump up ^ Uner, D. O. (1998). "A sensible mechanism of alkali promotion in Fischer Tropsch
synthesis:Adsorbate mobilities". Industrial and Engineering Chemistry Research 37 (6): 2239–2245.
doi:10.1021/ie970696d.
16. Jump up ^ Zupanc, C.; Hornung, A.; Hinrichsen, O.; Muhler, M. (2002). "The Interaction of Hydrogen
with Ru/MgO Catalysts". Journal of Catalysis 209 (2): 501–514. doi:10.1006/jcat.2002.3647.
17. Jump up ^ Trens, P.; Durand, R.; Coq, B.; Coutanceau, C.; Rousseau, S.; Lamy, C. (2009). "Poisoning of
Pt/C catalysts by CO and its consequences over the kinetics of hydrogen chemisorption". Applied Catalysis
B: Environmental 92 (3–4): 280–284. doi:10.1016/j.apcatb.2009.08.004.
18. Jump up ^ Rozanov, V. V.; Krylov, O. V. (1997). "Hydrogen spillover in heterogeneous catalysis".
Russian Chemical Reviews 66 (2): 107–119. doi:10.1070/RC1997v066n02ABEH000308.
19. Jump up ^ The Thermodynamics of Tetiris, Ars Technica, 2009.
20. Jump up ^ Barnes, Brian C.; Siderius, Daniel W.; Gelb, Lev D. (2009). "Structure, Thermodynamics, and
Solubility in Tetromino Fluids". Langmuir 25 (12): 6702–16. doi:10.1021/la900196b. PMID 19397254.
Cracking

http://www.eia.gov/todayinenergy/detail.cfm?id=9650

Introduction

Oil refinery in California. Photo: CA Energy Commission

Cracking is a petroleum refining process in which heavy-molecular weight hydrocarbons are


broken up into light hydrocarbon molecules by the application of heat and pressure, with or
without the use of catalysts, to derive a variety of fuel products. Cracking is one of the principal
ways in which crude oil is converted into useful fuels such as motor gasoline, jet fuel, and home
heatng oil.

Thermal Cracking
Thermal cracking is a refining process in which heat (~800°C) and pressure (~700kPa) are used
to break down, rearrange, or combine hydrocarbon molecules. The first thermal cracking process
was developed around 1913. Distillate fuels and heavy oils were heated under pressure in large
drums until they cracked into smaller molecules with better antiknock characteristics. However,
this method produced large amounts of solid, unwanted coke. This early process has evolved into
the following applications of thermal cracking: visbreaking, steam cracking, and coking.

Steam Cracking Process

Steam cracking is a petrochemical process sometimes used in refineries to produce olefinic raw
materials (e.g., ethylene) from various feedstock for petrochemicals manufacture. The feedstock
range from ethane to vacuum gas oil, with heavier feeds giving higher yields of by-products such
as naphtha. The most common feeds are ethane, butane, and naphtha. Steam cracking is carried
out at temperatures of 1,500°-1,600° F, and at pressures slightly above atmospheric. Naphtha
produced from steam cracking contains benzene, which is extracted prior to hydrotreating.
Residual from steam cracking is sometimes blended into heavy fuels.

Visbreaking Process

Visbreaking, a mild form of thermal cracking, significantly lowers the viscosity of heavy crude-
oil residue without affecting the boiling point range. Residual from the atmospheric distillation
tower is heated (800°-950° F) at atmospheric pressure and mildly cracked in a heater. It is then
quenched with cool gas oil to control overcracking, and flashed in a distillation tower.
Visbreaking is used to reduce the pour point of waxy residues and reduce the viscosity of
residues used for blending with lighter fuel oils. Middle distillates may also be produced,
depending on product demand. The thermally cracked residue tar, which accumulates in the
bottom of the fractionation tower, is vacuum flashed in a stripper and the distillate recycled.
(Source Mbeychok, via
Wikipedia.org)
http://en.wikipedia.org/wiki/File:Visbreaker.png#filelinks

Coking Processes

Coking is a severe method of thermal cracking used to upgrade heavy residuals into lighter
products or distillates. Coking produces straight-run gasoline (coker naphtha) and various
middle-distillate fractions used as catalytic cracking feedstock. The process so completely
reduces hydrogen that the residue is a form of carbon called "coke." The two most common
processes are delayed coking and continuous (contact or fluid) coking. Three typical types of
coke are obtained (sponge coke, honeycomb coke, and needle coke) depending upon the reaction
mechanism, time, temperature, and the crude feedstock.

Delayed Coking

In delayed coking the heated charge (typically residuum from atmospheric distillation towers) is
transferred to large coke drums which provide the long residence time needed to allow the
cracking reactions to proceed to completion. Initially the heavy feedstock is fed to a furnace
which heats the residuum to high temperatures (900°-950° F) at low pressures (25-30 psi) and is
designed and controlled to prevent premature coking in the heater tubes. The mixture is passed
from the heater to one or more coker drums where the hot material is held approximately 24
hours (delayed) at pressures of 25-75 psi, until it cracks into lighter products. Vapors from the
drums are returned to a fractionator where gas, naphtha, and gas oils are separated out. The
heavier hydrocarbons produced in the fractionator are recycled through the furnace.
http://en.wikipedia.org/wiki/File:Delayed_Coker.png
http://en.wikipedia.org/wiki/File:Delayed_Coker.png

After the coke reaches a predetermined level in one drum, the flow is diverted to another drum to
maintain continuous operation. The full drum is steamed to strip out uncracked hydrocarbons,
cooled by water injection, and decoked by mechanical or hydraulic methods. The coke is
mechanically removed by an auger rising from the bottom of the drum. Hydraulic decoking
consists of fracturing the coke bed with high-pressure water ejected from a rotating cutter.

Continuous Coking

Continuous (contact or fluid) coking is a moving-bed process that operates at temperatures


higher than delayed coking. In continuous coking, thermal cracking occurs by using heat
transferred from hot, recycled coke particles to feedstock in a radial mixer, called a reactor, at a
pressure of 50 psi. Gases and vapors are taken from the reactor, quenched to stop any further
reaction, and fractionated. The reacted coke enters a surge drum and is lifted to a feeder and
classifier where the larger coke particles are removed as product. The remaining coke is dropped
into the preheater for recycling with feedstock. Coking occurs both in the reactor and in the surge
drum. The process is automatic in that there is a continuous flow of coke and feedstock.
Catalytic Cracking
Catalytic cracking breaks complex hydrocarbons into simpler molecules in order to increase the
quality and quantity of lighter, more desirable products and decrease the amount of residuals.
This process rearranges the molecular structure of hydrocarbon compounds to convert heavy
hydrocarbon feedstock into lighter fractions such as kerosene, gasoline, liquified petroleum gas
(LPG), heating oil, and petrochemical feedstock.

Catalytic cracking is similar to thermal cracking except that catalysts facilitate the conversion of
the heavier molecules into lighter products. Use of a catalyst (a material that assists a chemical
reaction but does not take part in it) in the cracking reaction increases the yield of improved-
quality products under much less severe operating conditions than in thermal cracking. Typical
temperatures are from 850°-950° F at much lower pressures of 10-20 psi. The catalysts used in
refinery cracking units are typically solid materials (zeolite, aluminum hydrosilicate, treated
bentonite clay, fuller's earth, bauxite, and silica-alumina) that come in the form of powders,
beads, pellets or shaped materials called extrudites.

There are three basic functions in the catalytic cracking process:

 Reaction: Feedstock reacts with catalyst and cracks into different hydrocarbons;
 Regeneration: Catalyst is reactivated by burning off coke; and
 Fractionation: Cracked hydrocarbon stream is separated into various products.

The three types of catalytic cracking processes are fluid catalytic cracking (FCC), moving-bed
catalytic cracking, and Thermofor catalytic cracking (TCC). The catalytic cracking process is
very flexible, and operating parameters can be adjusted to meet changing product demand. In
addition to cracking, catalytic activities include dehydrogenation, hydrogenation, and
isomerization.

Fluid Catalytic Cracking (FCC)

See main article: Fluid catalytic cracking

Fluid catalytic cracking or "cat cracking," is the basic gasoline-making process. Using intense
heat (about 1,000 degrees Fahrenheit), low pressure and a powdered catalyst (a substance that
accelerates chemical reactions), the cat cracker can convert most relatively heavy fractions into
smaller gasoline molecules. The fluid cracker consists of a catalyst section and a fractionating
section that operate together as an integrated processing unit . The catalyst section contains the
reactor and regenerator, which, with the standpipe and riser, forms the catalyst circulation unit.
The fluid catalyst is continuously circulated between the reactor and the regenerator using air, oil
vapors, and steam as the conveying media.
https://www.isa.org/InTechTemplate.cfm?Section=Features3&template=/TaggedPage/DetailDisplay.cfm&ContentID=44006
https://www.isa.org/InTechTemplate.cfm?Section=Features3&template=/TaggedPage/DetailDisplay.cfm&ContentID=44006

A typical FCC process involves mixing a preheated hydrocarbon charge with hot, regenerated
catalyst as it enters the riser leading to the reactor. The charge is combined with a recycle stream
within the riser, vaporized, and raised to reactor temperature (900°-1,000° F) by the hot catalyst.
As the mixture travels up the riser, the charge is cracked at 10-30 psi. In the more modern FCC
units, all cracking takes place in the riser. The "reactor" no longer functions as a reactor; it
merely serves as a holding vessel for the cyclones. This cracking continues until the oil vapors
are separated from the catalyst in the reactor cyclones. The resultant product stream (cracked
product) is then charged to a fractionating column where it is separated into fractions, and some
of the heavy oil is recycled to the riser.

Spent catalyst is regenerated to get rid of coke that collects on the catalyst during the process.
Spent catalyst flows through the catalyst stripper to the regenerator, where most of the coke
deposits burn off at the bottom where preheated air and spent catalyst are mixed. Fresh catalyst is
added and worn-out catalyst removed to optimize the cracking process.

Moving Bed Catalytic Cracking

The moving-bed catalytic cracking process is similar to the FCC process. The catalyst is in the
form of pellets that are moved continuously to the top of the unit by conveyor or pneumatic lift
tubes to a storage hopper, then flow downward by gravity through the reactor, and finally to a
regenerator. The regenerator and hopper are isolated from the reactor by steam seals. The
cracked product is separated into recycle gas, oil, clarified oil, distillate, naphtha, and wet gas.

Thermofor Catalytic Cracking

In a typical thermofor catalytic cracking unit, the preheated feedstock flows by gravity through
the catalytic reactor bed. The vapors are separated from the catalyst and sent to a fractionating
tower. The spent catalyst is regenerated, cooled, and recycled. The flue gas from regeneration is
sent to a carbon monoxide boiler for heat recovery.

Hydrocracking
Hydrocracking is a two-stage process combining catalytic cracking and hydrogenation, wherein
heavier feedstocks are cracked in the presence of hydrogen to produce more desirable products.
The process employs high pressure, high temperature, a catalyst, and hydrogen. Hydrocracking is
used for feedstocks that are difficult to process by either catalytic cracking or reforming, since
these feedstocks are characterized usually by a high polycyclic aromatic content and/or high
concentrations of the two principal catalyst poisons, sulfur and nitrogen compounds.

The hydrocracking process largely depends on the nature of the feedstock and the relative rates
of the two competing reactions, hydrogenation and cracking. Heavy aromatic feedstock is
converted into lighter products under a wide range of very high pressures (1,000-2,000 psi) and
fairly high temperatures (750°-1,500° F), in the presence of hydrogen and special catalysts.
When the feedstock has a high paraffinic content, the primary function of hydrogen is to prevent
the formation of polycyclic aromatic compounds. Another important role of hydrogen in the
hydrocracking process is to reduce tar formation and prevent buildup of coke on the catalyst.
Hydrogenation also serves to convert sulfur and nitrogen compounds present in the feedstock to
hydrogen sulfide and ammonia.

Hydrocracking produces relatively large amounts of isobutane for alkylation feedstock.


Hydrocracking also performs isomerization for pour-point control and smoke-point control, both
of which are important in high-quality jet fuel.

Hydrocracking Process

In the first stage, preheated feedstock is mixed with recycled hydrogen and sent to the first-stage
reactor, where catalysts convert sulfur and nitrogen compounds to hydrogen sulfide and
ammonia. Limited hydrocracking also occurs.
http://www.eia.gov/todayinenergy/detail.cfm?id=9650
http://www.eia.gov/todayinenergy/detail.cfm?id=9650

After the hydrocarbon leaves the first stage, it is cooled and liquefied and run through a
hydrocarbon separator. The hydrogen is recycled to the feedstock. The liquid is charged to a
fractionator. Depending on the products desired (gasoline components, jet fuel, and gas oil), the
fractionator is run to cut out some portion of the first stage reactor out-turn. Kerosene-range
material can be taken as a separate side-draw product or included in the fractionator bottoms
with the gas oil.

The fractionator bottoms are again mixed with a hydrogen stream and charged to the second
stage. Since this material has already been subjected to some hydrogenation, cracking, and
reforming in the first stage, the operations of the second stage are more severe (higher
temperatures and pressures). Like the outturn of the first stage, the second stage product is
separated from the hydrogen and charged to the fractionator.

Catalytic Reforming
Catalytic reforming is an important process used to convert low-octane naphthas into high-octane
gasoline blending components called reformates. Reforming represents the total effect of
numerous reactions such as cracking, polymerization, dehydrogenation, and isomerization taking
place simultaneously. Depending on the properties of the naphtha feedstock (as measured by the
paraffin, olefin, naphthene, and aromatic content) and catalysts used, reformates can be produced
with very high concentrations of toluene, benzene, xylene, and other aromatics useful in gasoline
blending and petrochemical processing. Hydrogen, a significant by-product, is separated from
the reformate for recycling and use in other processes.
http://en.wikipedia.org/wiki/File:CatReformer.png
http://en.wikipedia.org/wiki/File:CatReformer.png

A catalytic reformer comprises a reactor section and a product-recovery section. More or less
standard is a feed preparation section in which, by combination of hydrotreatment and
distillation, the feedstock is prepared to specification. Most processes use platinum as the active
catalyst. Sometimes platinum is combined with a second catalyst (bimetallic catalyst) such as
rhenium or another noble metal.

There are many different commercial catalytic reforming processes including platforming,
powerforming, ultraforming, and Thermofor catalytic reforming. In the platforming process, the
first step is preparation of the naphtha feed to remove impurities from the naphtha and reduce
catalyst degradation. The naphtha feedstock is then mixed with hydrogen, vaporized, and passed
through a series of alternating furnace and fixed-bed reactors containing a platinum catalyst. The
effluent from the last reactor is cooled and sent to a separator to permit removal of the hydrogen-
rich gas stream from the top of the separator for recycling. The liquid product from the bottom of
the separator is sent to a fractionator called a stabilizer (butanizer). It makes a bottom product
called reformate; butanes and lighter go overhead and are sent to the saturated gas plant.

Some catalytic reformers operate at low pressure (50-200 psi), and others operate at high
pressures (up to 1,000 psi). Some catalytic reforming systems continuously regenerate the
catalyst in other systems. One reactor at a time is taken off-stream for catalyst regeneration, and
some facilities regenerate all of the reactors during turnarounds.

Catalytic hydrotreating
Catalytic hydrotreating is a hydrogenation process used to remove about 90% of contaminants
such as nitrogen, sulfur, oxygen, and metals from liquid petroleum fractions. These
contaminants, if not removed from the petroleum fractions as they travel through the refinery
processing units, can have detrimental effects on the equipment, the catalysts, and the quality of
the finished product. Typically, hydrotreating is done prior to processes such as catalytic
reforming so that the catalyst is not contaminated by untreated feedstock. Hydrotreating is also
used prior to catalytic cracking to reduce sulfur and improve product yields, and to upgrade
middle-distillate petroleum fractions into finished kerosene, diesel fuel, and heating fuel oils. In
addition, hydrotreating converts olefins and aromatics to saturated compounds.

Catalytic Hydrodesulfurization Process

Hydrotreating for sulfur removal is called hydrodesulfurization. In a typical catalytic


hydrodesulfurization unit, the feedstock is deaerated and mixed with hydrogen, preheated in a
fired heater (600°-800° F) and then charged under pressure (up to 1,000 psi) through a fixed-bed
catalytic reactor. In the reactor, the sulfur and nitrogen compounds in the feedstock are converted
into hydrogen sulfide (H2S) and ammonia (NH3). The reaction products leave the reactor and
after cooling to a low temperature enter a liquid/gas separator. The hydrogen-rich gas from the
high-pressure separation is recycled to combine with the feedstock, and the low-pressure gas
stream rich in H2S is sent to a gas treating unit where H2S is removed. The clean gas is then
suitable as fuel for the refinery furnaces. The liquid stream is the product from hydrotreating and
is normally sent to a stripping column for removal of H2S and other undesirable components. In
cases where steam is used for stripping, the product is sent to a vacuum drier for removal of
water. Hydrodesulfurized products are blended or used as catalytic reforming feedstock.

Other Hydrotreating Processes

Hydrotreating processes differ depending upon the feedstock available and catalysts used.
Hydrotreating can be used to improve the burning characteristics of distillates such as kerosene.
Hydrotreatment of a kerosene fraction can convert aromatics into naphthenes, which are cleaner-
burning compounds.

Lube-oil hydrotreating uses catalytic treatment of the oil with hydrogen to improve product
quality. The objectives in mild lube hydrotreating include saturation of olefins and improvements
in color, odor, and acid nature of the oil. Mild lube hydrotreating also may be used following
solvent processing. Operating temperatures are usually below 600° F and operating pressures
below 800 psi. Severe lube hydrotreating, at temperatures in the 600°-750° F range and hydrogen
pressures up to 3,000 psi, is capable of saturating aromatic rings, along with sulfur and nitrogen
removal, to impart specific properties not achieved at mild conditions.

Hydrotreating also can be employed to improve the quality of pyrolysis gasoline (pygas), a by-
product from the manufacture of ethylene. Traditionally, the outlet for pygas has been motor
gasoline blending, a suitable route in view of its high octane number. However, only small
portions can be blended untreated owing to the unacceptable odor, color, and gum-forming
tendencies of this material. The quality of pygas, which is high in diolefin content, can be
satisfactorily improved by hydrotreating, whereby conversion of diolefins into mono-olefins
provides an acceptable product for motor gas blending.

Вам также может понравиться