catalysis

catalysis
catalytic /kat'l it"ik/, adj., n.catalytical, adj.catalytically, adv.
/keuh tal"euh sis/, n., pl. catalyses /-seez'/.
1. Chem. the causing or accelerating of a chemical change by the addition of a catalyst.
2. an action between two or more persons or forces, initiated by an agent that itself remains unaffected by the action: social catalyses occasioned by controversial writings.
[1645-55; < NL < Gk katálysis dissolution, equiv. to katalý(ein) to dissolve (kata- CATA- + lýein to loosen) + -sis -SIS]

* * *

Modification (usually acceleration) of a chemical reaction rate by addition of a catalyst, which combines with the reactants but is ultimately regenerated so that its amount remains unchanged and the chemical equilibrium of the conditions of the reaction is not altered.

Catalysts reduce the activation energy barrier between reactants and products. When more than one reaction is possible, a catalyst that accelerates only one reaction pathway selectively enhances the creation of its product. Catalysis is inhibited if the reactant or the catalyst is removed or altered by any of several types of agents (inhibitors). Catalysis in a single phase (e.g., the catalyst is dispersed in a liquid solution or gaseous mixture with the reactants) is homogeneous; that in more than one phase (e.g., the reactants are liquids and the catalyst a solid) is heterogeneous. Chemisorption, a type of heterogeneous catalysis, often involves bonding between the catalyst's solid surface and the reactant, changing the nature of the chemisorbed molecules. To make the accessible surface area as large as possible, such catalysts are finely powdered or highly porous solids. Catalysis is essential to the modern chemical industry. See also enzyme.

* * *

▪ chemical process
Introduction

      in chemistry, the modification of the rate of a chemical reaction, usually an acceleration, by addition of a substance not consumed during the reaction. The rates of chemical reactions—that is, the velocities at which they occur—depend upon a number of factors, including the chemical nature of the reacting species and the external conditions to which they are exposed. A particular phenomenon associated with the rates of chemical reactions that is of great theoretical and practical interest is catalysis, the acceleration of chemical reactions by substances not consumed in the reactions themselves—substances known as catalysts (catalyst). The study of catalysis is of interest theoretically because of what it reveals about the fundamental nature of chemical reactions; in practice, the study of catalysis is important because many industrial processes depend upon catalysts for their success. Fundamentally, the peculiar phenomenon of life would hardly be possible without the biological catalysts termed enzymes (enzyme).

      In a catalyzed reaction, the catalyst generally enters into chemical combination with the reactants but is ultimately regenerated, so the amount of catalyst remains unchanged. Since the catalyst is not consumed, each catalyst molecule may induce the transformation of many molecules of reactant. For an active catalyst, the number of molecules transformed per minute by one molecule of catalyst may be as large as several million.

      Where a given substance or a combination of substances undergoes two or more simultaneous reactions that yield different products, the distribution of products may be influenced by the use of a catalyst that selectively accelerates one reaction relative to the other(s). By choosing the appropriate catalyst, a particular reaction can be made to occur to the extent of practically excluding another. Many important applications of catalysis are based on selectivity of this kind.

      Since a reverse chemical reaction may proceed by reversal of the steps constituting the mechanism of the forward reaction, the catalyst for a given reaction accelerates the reaction in both directions equally. Therefore, a catalyst does not affect the position of equilibrium of a chemical reaction; it affects only the rate at which equilibrium is attained. Apparent exceptions to this generalization are those reactions in which one of the products is also a catalyst for the reaction. Such reactions are termed autocatalytic.

      Cases are also known in which the addition of a foreign substance, called an inhibitor, decreases the rate of a chemical reaction. This phenomenon, properly termed inhibition or retardation, is sometimes called negative catalysis. Concentrations of the inhibitor may in some cases be much lower than those of the reactants. Inhibition may result from (1) a decrease in the concentration of one of the reactants because of complex formation between the reactant and the inhibitor, (2) a decrease in the concentration of an active catalyst (“poisoning” of the catalyst) because of complex formation between the catalyst and the inhibitor, or (3) a termination of a chain reaction because of destruction of the chain carriers by the inhibitor.

History
      The term catalysis (from the Greek kata-, “down,” and lyein, “loosen”) was first employed by the great Swedish chemist Jöns Jacob Berzelius (Berzelius, Jöns Jacob) in 1835 to correlate a group of observations made by other chemists in the late 18th and early 19th centuries. These included the enhanced conversion of starch to sugar by acids (acid) first observed by Gottlieb Sigismund Constantin Kirchhoff; Sir Humphry Davy's (Davy, Sir Humphry, Baronet) observations that platinum hastens the combustion of a variety of gases (gas); the discovery of the stability of hydrogen peroxide in acid solution but its decomposition in the presence of alkali and such metals (metal) as manganese, silver, platinum, and gold; and the observation that the oxidation of alcohol to acetic acid is accomplished in the presence of finely divided platinum. The agents promoting these various reactions were termed catalysts (catalyst), and Berzelius postulated a special unknown catalytic force to be operating in such processes.

      In 1834 the English scientist Michael Faraday (Faraday, Michael) had examined the power of a platinum plate to accomplish the recombination of gaseous hydrogen and oxygen (the products of electrolysis of water) and the retardation of that recombination by the presence of other gases, such as ethylene and carbon monoxide. Faraday maintained that essential for activity was a perfectly clean metallic surface (at which the retarding gases could compete with the reacting gases and so suppress activity), a concept that would later be shown to be generally important in catalysis.

      Many of the primitive technical arts involved unconscious applications of catalysis. The fermentation of wine to acetic acid and the manufacture of soap (soap and detergent) from fats (fat) and alkalies were well known in man's early history. Sulfuric acid prepared by firing mixtures of sulfur and nitre ( sodium nitrate) was an early forerunner of the lead chamber process of sulfuric acid manufacture, in which sulfur dioxide oxidation was accelerated by the addition of oxides (oxide) of nitrogen. (A mechanism for the latter process was suggested by Sir Humphry Davy in 1812 on the basis of experiments carried out by others.)

      In 1850 the concept of a velocity of reaction was developed during studies of hydrolysis, or inversion, of cane sugar. The term inversion refers to the change in rotation undergone by monochromatic light when it is passed through the reaction system, a parameter that is easily measured, thereby facilitating study of the reaction. It was found that, at any moment, the rate of inversion was proportional to the amount of cane sugar undergoing transformation and that the rate was accelerated by the presence of acids. (Later it was shown that the rate of inversion was directly proportional to the strength of the acid.) This work was in part the precursor of later studies of reaction velocity and the accelerating influence of higher temperature on that velocity by J.H. van 't Hoff (Hoff, Jacobus Henricus van 't), Svante Arrhenius (Arrhenius, Svante August), and Wilhelm Ostwald (Ostwald, Wilhelm), all of whom played leading roles in the developing science of physical chemistry (chemistry). Ostwald's work on reaction velocities led him in the 1890s to define catalysts as substances that change the velocity of a given chemical reaction without modification of the energy factors of the reaction.

      This statement of Ostwald was a memorable advance since it implied that catalysts do not change the position of equilibrium in a reaction. In 1877 Georges Lemoine had shown that the decomposition of hydriodic acid to hydrogen and iodine reached the same equilibrium point at 350 °C (660 °F), 19 percent, whether the reaction was carried out rapidly in the presence of platinum sponge or slowly in the gas phase. This observation has an important consequence: a catalyst for the forward process in a reaction is also a catalyst for the reverse reaction. P.E.M. Berthelot (Berthelot, Pierre-Eugène-Marcellin), the distinguished French chemist, confirmed this observation in 1879 with liquid systems, when he found that the reaction of organic acids and alcohols, called esterification, is catalyzed by the presence of small amounts of a strong inorganic acid, just as is the reverse process, the hydrolysis of esters (ester) (the reaction between an ester and water).

      The deliberate application of catalysts to industrial processes was undertaken in the 19th century. P. Phillips, an English chemist, patented the use of platinum to oxidize sulfur dioxide to sulfur trioxide with air. His process was employed for a time but was abandoned because of loss of activity by the platinum catalyst. Poisons in the reactants were subsequently found to be responsible, and the process became a technical success at the turn of the 20th century. In 1871 an industrial process was developed for the oxidation of hydrochloric acid to chlorine in the presence of cupric salts impregnated in clay brick. The chlorine obtained was employed in the manufacture of bleaching powder (a dry substance that releases chlorine on treatment with acid) by reaction with lime. Again, in this reaction, it was observed that the same equilibrium was reached in both directions. Furthermore, it was found that the lower the temperature, the greater the equilibrium content of chlorine; a working temperature of 450 °C (840 °F) produced the maximum amount of chlorine in a convenient time.

      Toward the close of the 19th century, the classic studies of the eminent French chemist Paul Sabatier (Sabatier, Paul) on the interaction of hydrogen with a wide variety of organic compounds (organic compound) were carried out using various metal catalysts; this research led to the development of a German patent for the hydrogenation of liquid unsaturated fats to solid saturated fats with nickel catalysts. The development of three important German catalytic processes had great impact on industry at the end of the 19th century and in the early decades of the 20th. One was the so-called contact process for producing sulfuric acid catalytically from the sulfur dioxide produced by smelting operations. Another was the catalytic method for the synthetic production of the valuable dyestuff indigo. The third was the catalytic combination of nitrogen and hydrogen for the production of ammonia—the Haber-Bosch process for nitrogen fixation—developed by the chemists Fritz Haber (Haber, Fritz) and Carl Bosch (Bosch, Carl).

Classification of catalysts
      Catalysts (catalyst) may be classified generally according to their physical state, their chemical nature, or the nature of the reactions that they catalyze.

      Catalysts may be gases (gas), liquids (liquid), or solids (solid). In homogeneous catalysis, the catalyst is molecularly dispersed in the same phase (usually gaseous or liquid) as the reactants. In heterogeneous catalysis the reactants and the catalyst are in different phases, separated by a phase boundary. Most commonly, heterogeneous catalysts are solids, and the reactants are gases or liquids.

Homogeneous catalysis
      When the catalyst and the reacting substances are present together in a single state of matter, usually as a gas or a liquid, it is customary to classify the reactions as cases of homogeneous catalysis. Oxides (oxide) of nitrogen serve as catalysts for the oxidation of sulfur dioxide in the lead chamber process for producing sulfuric acid, an instance of homogeneous catalysis in which the catalyst and reactants are gases. Traces of water vapour catalyze some gas reactions—for example, the interaction of carbon monoxide and oxygen, which proceeds only slowly in dry conditions. Sulfuric acid used as a catalyst for the formation of diethyl ether from ethyl alcohol is an example of homogeneous catalysis in the liquid phase (when the products, water and ether, are continuously removed by distillation); by this method, considerable quantities of alcohol can be converted to ether with a single charge of sulfuric acid. The inversion of cane sugar and the hydrolysis of esters (ester) by acid solutions also are examples of homogeneous catalysis in the liquid phase.

      The oxidation of sodium sulfite solutions by dissolved oxygen is greatly accelerated by minute traces of copper ions (ion) in the homogeneous liquid system. This system is of special interest since it has been shown that the process is a chain reaction. In this case many thousands of molecules (molecule) of sodium sulfite can be oxidized to sulfate if the initial activation process is produced by absorption of a limited number of quanta (quantum) (discrete energy measures) of light. The best example of a light-initiated chain reaction is the photocombination of hydrogen (H2) and chlorine (Cl2); as many as one million molecules of hydrogen chloride can be formed by absorption of a single light quantum (designated hν). Here the sequence of reaction is

      with reactions 2 and 3 repeated over and over again. It is of interest to note that such chain reactions can be retarded by the presence of negative catalysts, more commonly termed inhibitors. These are materials that slow down the overall reaction by shortening the reaction chains, generally by entering into a non-chain reaction with one of the chemical components that maintain the chain. A wide variety of substances—including alcohols, sugars, and phenols (phenol)—have been found to act as inhibitors of the oxidation of sulfite solutions.

      A generalized treatment of homogeneous catalysis by acids and bases (base) was given by the Danish physical chemist J.N. Brønsted (Brønsted, Johannes Nicolaus) in the mid-1920s on the basis of his concept of acids and bases. According to Brønsted, an acid is a molecule that can furnish a proton, and a base is a molecule that takes up a proton. On this assumption, the range of acids includes such varied materials as bisulfate ion, HSO4; acetic acid, CH3COOH; water, H2O; hydronium ion, H3O+; and ammonium ion, NH4+. The corresponding bases are sulfate ion, SO42−; acetate ion, CH3COO; hydroxide ion, OH; water, H2O; and ammonia, NH3 (these substances accept protons to yield the listed acids). Brønsted studied a number of acid- and base-catalyzed reactions, including (1) the acid-catalyzed hydrolysis of an ester, ethyl orthoacetate, (2) the basic catalysis of nitramide decomposition (H2N–NO2→ H2O + N2O), and (3) the acid-base catalysis of the conversion (mutarotation) of glucose to a closely related form. In each case, he observed a direct relationship between the velocity of the catalyzed reaction and the concentration of the catalyzing substance.

      Based in part on the ideas of Brønsted, a general scheme for a change of a substance A to another substance B catalyzed by a material C can be formulated thus:

      The designation Z refers to an intermediate stage, which is formed with a velocity indicated by k1; Z can disappear either to reform A + C, with a velocity k2, or it can decompose by path 2, with velocity k3, to give the product B and regenerate the catalyst C. If k3 is much greater than k2, the intermediate Z is used up almost as quickly as it is formed. The product will then be formed at a rate governed by the expression k1[A][C], in which the square brackets indicate concentrations of reactant and catalyst. If k2 is much larger than k3, however, the velocity-determining process is the decomposition of Z, the rate of formation of the product being represented by the expression k3[Z], in which [Z] is the concentration of the intermediate. Examples of both types of change have been studied.

      In certain instances two or more catalysts present at the same time produce effects greater than either would produce alone. It is then customary to speak of promoter action. Thus, iron ions in solution fortify the action of copper ions in catalyzing a reaction between hydrogen peroxide and iodine. It is assumed that each catalyst activates only one of the reactants.

      The most important modern examples of homogeneous catalyses are found in the petrochemical industry. The oxo reaction is one such process: carbon monoxide and hydrogen are added to olefins (olefin) (unsaturated hydrocarbons (hydrocarbon)) at around 150 °C (300 °F) and 200 atmospheres of pressure to form aldehydes (aldehyde) and alcohols, oxygen-containing organic compounds. A cobalt carbonyl catalyst Co2(CO)8 is employed; this hydrocarbon-soluble catalyst is believed to activate hydrogen by formation of HCo(CO)4, which then reacts with the olefin. This reaction has led to a number of studies of organometallic chemistry. copper, silver, and mercury cations (cation) (positively charged ions (ion)) and permanganate anions (anion) (negative ions) also are known to act as homogeneous catalysts for hydrogen activation. palladium chloride is employed industrially in the catalytic oxidation of ethylene to acetaldehyde in the presence of cupric chloride. The palladium is presumed to be repeatedly converted from the salt to the free metal, the function of the cupric chloride being to participate in the re-formation of the palladium salt from the metal.

      Phosphoric (phosphoric acid), sulfuric, sulfonic (sulfonic acid), and hydrobromic acids are important agents in the industrial processes of isomerization, polymerization, hydration, and dehydration, as well as in the classic esterification reactions. Free radicals (molecular fragments bearing unpaired electrons (electron)) that are generated by the decomposition of peroxides (peroxide) or metal alkyls also initiate homogeneous catalytic processes.

Heterogeneous catalysis
      Many catalytic processes are known in which the catalyst and the reactants are not present in the same phase—that is, state of matter. These are known as heterogeneous catalytic reactions. They include reactions between gases or liquids or both at the surface of a solid catalyst. Since the surface is the place at which the reaction occurs, it generally is prepared in ways that produce large surface areas per unit of catalyst; finely divided metals, metal gauzes, metals incorporated into supporting matrices, and metallic films have all been used in modern heterogeneous catalysis. The metals themselves are used, or they are converted to oxides, sulfides (sulfide), or halides (halide mineral).

 With solid catalysts, at least one of the reactants is chemisorbed (a portmanteau term for chemically adsorbed) by the catalyst. A catalyst is chosen that releases the products formed as readily as possible; otherwise the products remain on the catalyst surface and act as poisons to the process. Chemisorption can occur over a wide temperature range, the most effective temperature for adsorption depending on the nature of the catalyst. Thus, hydrogen is chemisorbed readily by many metals even at liquid air temperatures (below −180 °C [−290 °F]). With a series of hydrogenation-dehydrogenation catalysts—e.g., zinc oxide–chromic oxide (ZnO–Cr2O3)—chemisorption of hydrogen often occurs above room temperature. Nitrogen is rapidly chemisorbed on synthetic ammonia- iron catalyst in the region above 400 °C (750 °F). It has been shown that iron films chemisorb nitrogen even at liquid air temperatures, with additional chemisorption found above room temperatures. It follows from such considerations that whereas physical adsorptions, which parallel the ease of liquefaction of the adsorbed substance, occur spontaneously, chemisorption, which involves the making and breaking of chemical bonds, often requires activation energies (activation energy) (energy needed to initiate reactions) as do uncatalyzed chemical processes. To be efficient catalytically, a process must involve energies of activation for all the steps involved that, at their maxima, are less than those required for the uncatalyzed reaction. This situation is illustrated graphically in the figure—>, with a hypothetical reaction that could occur by either an uncatalyzed or a catalyzed route.

      Two competing proposals have been made concerning the mechanism of catalytic reactions at surfaces, and it has not been possible to choose between them. Originally, Irving Langmuir (Langmuir, Irving), an American physical chemist, proposed chemisorption of both reacting species at the surface, followed by interaction between adjacent species and evaporation of the products. An alternative proposal involves interaction between an impinging molecule and species already adsorbed on the surface. Subsequent developments have suggested various modes of attachment of the adsorbed and adsorbing species.

      A major advance in the science of surface catalysis was the development of a method for determining the surface area of catalysts (and other materials) by measuring the multimolecular adsorption of nitrogen at liquid nitrogen temperatures or the adsorption of other gases close to their boiling points (boiling point). It then became possible to calculate a quantity, designated Vm, that represents the volume of gas necessary to form a monolayer on the accessible surface; furthermore, the area of the surface can be determined from the known dimensions of the adsorbed molecules. It has also been found possible to titrate (measure quantitatively) the area of surfaces by chemisorption of gases. Since heterogeneously catalyzed reactions occur on the surface of the catalyst, the rates of such reactions are proportional to the accessible surface area of the catalyst. Active catalysts are thus usually highly porous solids with total surface areas as high as several hundred square metres per gram.

      When measurements of surface areas became possible, it was seen at once that many constituents present in minor quantities in the main catalyst material—known as promoters—could act by extending the effective surface area of the catalyst. It also was shown, however, that a promoter might produce an increase in the quality of the surface for the given reaction. Acting in a reverse direction are minor constituents of the reacting system or unwanted products of the reaction, which by preferential adsorption on the reaction sites and resistance to removal give rise to poisons for the process. Poisoning of a catalyst may also result from the poison adversely modifying the electronic properties of the catalyst.

      Much can be learned about mechanisms of surface processes by studying the behaviour of isotopic species of the reactants and products on the catalyst. An example of such use concerns the technically important synthesis of ammonia from its elements, the well-known Haber-Bosch process on promoted-iron catalysts. The synthesis of ammonia involves three types of bonds—hydrogen-hydrogen, nitrogen-hydrogen, and nitrogen-nitrogen—all of which can be studied using isotopes (isotope) of hydrogen and nitrogen. The first of these can be examined on the reduced-iron catalyst by following the progress of the reaction H2 + D2→ 2HD (in which D equals the deuterium atom, an isotope of hydrogen) on the surface. The reaction is found to occur rapidly even at liquid air temperatures. Nitrogen-hydrogen bond activation can be studied by following the reaction NH3 + ND3→ NH2D + NHD2. This proceeds steadily at room temperatures. The reaction involving only N–N bonds, however, studied by following the process 14N2 + 15N2→ 214N–15N (in which 14N and 15N are stable isotopes of nitrogen), is shown to proceed only at the higher temperatures of ammonia synthesis, around 400 °C (750 °F). From these data one concludes that the activation of the nitrogen molecules is the slow step (the process that limits the overall reaction) in ammonia synthesis. This conclusion is confirmed by measurement of rate of adsorption of nitrogen on the iron catalysts. Other, similar isotopic studies have yielded valuable information on the reactions of hydrocarbons, using deuterium and carbon-14 as the isotopic tracers.

Catalysis in stereoregular polymerization
      The importance of the concept of adsorption of reactants on the surface of catalysts (catalyst) has been greatly increased by the development of stereoregular polymerization processes—that is, methods that yield polymers whose molecules (molecule) have definite three-dimensional patterns. Such processes were developed independently by the German chemist Karl Ziegler (Ziegler, Karl) and the Italian Giulio Natta (Natta, Giulio). An example is the polymerization of propylene with a titanium trichloride–alkyl aluminum catalyst. In the case of a generalized ethylenic compound, CH2=CHR, stereoregular polymerization may yield three different arrangements of the polymer: an isotactic polymer, a syndiotactic polymer, and an atactic polymer. These have the following arrangements of their molecular chains:

      In the isotactic polymer the monomer units have added head-to-tail, to give a series of C–R tertiary bonds with the same configuration in space; in the syndiotactic polymer the tertiary carbon atoms in the chain have alternate (dextro and levo) spatial configurations; and in the atactic polymer there is no regularity in the distribution of steric configurations of the asymmetric carbon atoms. The various polymeric forms differ in their physical properties. Isotactic polypropylene, for example, has a density of 0.92 gram per cubic cm (0.53 ounce per cubic inch) and a melting point of 165 °C (329 °F), whereas an atactic polymer has a somewhat smaller density, 0.85 gram per cubic cm (0.49 ounce per cubic inch), and a much lower melting point of −35 °C (−30 °F). The more regular isotactic polymer is denser and has a higher melting point than the atactic product because of its greater tendency to crystallize (in spite of the fact that the substituent R may be quite large, hindering crystal formation). Stereoregular polymerization suggests a stereoregulated adsorption at the active centres of the catalyst. In the case of polypropylene, the catalytic centres have been identified by electron micrographs as α-TiCl3 surfaces, which cover only a small fraction of the total surface area, whereas the β-TiCl3 surfaces, which are more abundant, appear to be covered with polymer. The difference between the α- and β-surfaces lies in the random (α) and linear (β) arrangements of Ti3+ sites in the two surfaces.

      Since the Ziegler-Natta studies, other stereoregulating catalysts have been investigated, notably oxides (oxide) of chromium, vanadium, molybdenum, and tungsten on silica- alumina or other supports. Other cationic, anionic, and free-radical catalysts are known to produce stereoregulated polymerization. Stereoregular polymerization of dienes has undergone industrial development with the polymerization of isoprene to synthetic natural rubber.

Determination of the structure and properties of catalysts
      The nature of the active centres in catalytic material is further demonstrated by the enhancement of the catalytic activity of relatively inactive materials when they are subjected to intense radiation. silica gel bombarded by gamma rays (gamma ray) from cobalt-60 turns purplish in colour and becomes capable of inducing the reaction H2 + D2→ 2HD at liquid- nitrogen temperatures. The colour centres, which are positive “holes” (deficiencies) trapped in the vicinity of an oxygen ion next to an aluminum impurity, are bleached in vacuo above 200 °C (400 °F) and are destroyed by hydrogen even at room temperature.

      The properties of dilute concentrations of platinum metals (metal) in oxide matrices, such as silica and alumina, as well as on carbon carriers have been studied by Russian and American scientists. Such catalysts (catalyst) have technical significance in processes for the reforming of gasoline. In such catalysts—containing about 0.5 percent by weight of platinum or palladium—the degree of dispersion of the metal (that is, the ratio of the number of surface metal atoms (atom) to the total number present) is close to one. By contrast, on platinum foil the dispersion is only about 4 × 10−3. The titration and adsorption procedures with hydrogen and oxygen are employed to evaluate these dispersions.

      From these studies it becomes clear that there are two types of behaviour resulting from dispersion. For numerous catalytic processes, ranging from hydrogen- deuterium exchange to the hydrogenation of benzene and the hydrogenolysis of cyclopentane, the reactions are independent of dispersion in the critical region—with catalyst particle size of 5 nm or less. Such structure-insensitive processes have been termed facile reactions. On the other hand, there are reactions such as the isomerization of neopentane to isopentane and simultaneous cracking of the latter to isobutane and methane on platinum-alumina catalysts, where the selectivity for isomerization varies by a factor of 100 for the various catalysts studied (when the hydrogen-neopentane ratio is 10). Thus, the same 1 percent platinum-on-carbon catalyst showed a selectivity ratio of isomerization to hydrogenolysis of 2.5 when the catalyst was reduced in hydrogen at 500 °C (900 °F) and a selectivity ratio of 13 when the catalyst was fired in vacuo at 900 °C (1,600 °F), the percentage dispersion remaining at 35 percent in both cases. Such structure-sensitive catalytic reactions have been called “demanding reactions.” The gain in selectivity appears to be largely because of a reduction in the rate of hydrogenolysis. Since other studies have shown that heating in vacuo to 900 °C tends to develop certain (111) facets of the metal, it is thought that the increase in selectivity is due to a more abundant triadsorption of neopentane on the samples fired at high temperature. It has been shown that a crystallite of platinum about 2 nm in size has unusual surfaces not present in a regular octahedral crystallite of similar size. A number of sites where an adsorbed molecule could be surrounded by five platinum nearest neighbours were found on the crystallite with the unusual surface.

      An alternative approach to the problem of surface catalysis involves the consideration of electronic factors in catalyst and reactants. Many catalytic materials are semiconductors (semiconductor). It is thought that these can form a variety of bonds with reactants depending on the free lattice electrons (electron) and the holes in the catalyst lattice. Chemisorbed particles react in ways that are dependent on the form of attachment to the surface and that vary with the extent of coverage of the surface as well as with the available supply of electrons and holes. The surface behaves as would free radicals that are introduced directly into the reacting species, dependent on the electrochemical properties of the surface and the bulk of the semiconductor material. Such considerations have led to the determination of the character of the catalyst as a semiconductor and of the adsorbate as an electrochemical species, whether it is composed of positive or negative ions (ion) or free atoms or radicals. Catalytic activity has also been explored as a function of the d-band character—that is, the number of electrons in d orbitals (orbital) in the atoms of the catalyst materials.

      Since 1940 various instrumental techniques have been developed to explore the structure of catalytic materials and the character of the adsorbed species, even during the reaction itself. Among these techniques are electron microscopy, field emission microscopy, electron microprobe methods, magnetic measurements, infrared spectroscopy, Mössbauer spectroscopy, measurements of heats of immersion, flash desorption procedures, low-energy electron diffraction studies, and nuclear magnetic resonance and electron spin resonance techniques.

Other catalytic compounds
      The term polyfunctional heterogeneous catalysis is applied to a group of catalysts (catalyst) in which more than one component of the surface is active in the processes under study. One example of a bifunctional heterogeneous catalyst is the catalyst of metal ( platinum or nickel) deposited on a silica- alumina “acidic” base. Such dual functional catalysts are involved in the interconversions of saturated hydrocarbons (hydrocarbon) (paraffins) and unsaturated hydrocarbons (olefins (olefin)) and normal (straight-chain) and iso (branched-chain) hydrocarbons, as well as in the splitting (cracking) of the hydrocarbon molecules (molecule). The interrelations involved are as follows:

      in which X1 and X2 are metal-catalyzed processes and Y1 and Y2 are acid-catalyzed processes. Operating conditions can be altered to maximize the hydrocracking reactions relative to hydrogenolysis.

      A variety of catalysts with “acidic” sites have been found to be active in the dehydration of alcohols (alcohol) and in the cracking and isomerization of hydrocarbons. Among these are silica, obtained by calcination (heating) of silica gel; high-purity alumina, prepared by the calcining of specially prepared aluminum hydroxide; and silica-alumina mixtures. The catalytic sites have been found to have varying degrees of acidity; their exact nature, as well as their characterization in terms of the atomic architecture of the solid catalyst, is still under discussion. In the case of silica-alumina, the sites are ascribed to the presence of trivalent aluminum ions (ion), Al3+, in a matrix of quadrivalent silicon ions, Si4+, which gives rise to charge differences in the neighbourhood of the aluminum ions. These acidic sites can be poisoned by ammonia and amines (amine), a finding that confirms their acidic nature. When these catalysts are treated with alkalies (alkali), their catalytic character is greatly modified. On the other hand, treatment with halogen elements (halogen element), especially fluorine and chlorine, enhances the acidic properties of these oxide materials.

      Zeolites (zeolite) are naturally occurring crystalline aluminosilicates that have a porous structure and contain cations (cation), generally of the alkali (alkali metal) or alkaline earth metals. The cations can be exchanged reversibly with other metal ions without destroying the aluminosilicate structure. Because the zeolites rapidly adsorb certain molecules and exclude others, they have been given the name “molecular sieves.” The adsorption characteristics of natural and synthetic zeolites have been studied since the 1930s. Manufactured zeolites, some of which have structures not found in nature, are employed as dehydrating agents but also may be used for the production of catalytic materials by exchange with cationic elements or by impregnation of metal salt solutions into the pores of the zeolite; a large number of zeolitic catalysts have been developed.

      A class of compounds termed electron donor-acceptor complexes also has been studied for its catalytic activity. The class may be exemplified by a complex between metallic sodium (the donor) and anthracene, C14H10, a tricyclic hydrocarbon (the acceptor). The complex can be visualized as an anthracene anion and a sodium cation. Such complexes can exchange the hydrogen of the anion with molecular hydrogen that has been brought into contact with the complex. A complex represented by ZH (in which Z represents all of the molecule except for the exchangeable hydrogen) could undergo an exchange with deuterium as follows: ZH + D2→ ZD + HD. It could also take part in corresponding exchanges with hydrocarbons or bring about hydrogenation of hydrocarbons. Among other electron-acceptor catalysts are the metal phthalocyanines (compounds related to certain biological catalysts) and activated charcoal. Some donor-acceptor complexes synthesize ammonia from nitrogen-hydrogen mixtures. This reaction represents a close approach to the activity of biological and bacterial catalysts.

Biological catalysts: the enzymes (enzyme)
      Enzymes (enzyme) are substances found in biological systems that are catalysts (catalyst) for specific biochemical processes. Although earlier discoveries of enzymes had been made, a significant confirmation of their importance in living systems was found in 1897 by the German chemist Eduard Buchner (Buchner, Eduard), who showed that the filtered cell-free liquor from crushed yeast cells could bring about the conversion of sugar to carbon dioxide. Since that time more than 1,000 enzymes have been recognized, each specific to a particular chemical reaction occurring in living systems. More than 100 of these have been isolated in relatively pure form, including a number of crystallized enzymes. The first enzymes to be crystallized were urease, isolated from the jack bean and crystallized in 1926 by James Batcheller Sumner (Sumner, James Batcheller), and pepsin, crystallized in 1930 by John Howard Northrop (Northrop, John Howard), both of whom won the Nobel Prize for Chemistry for their work. These purified materials were shown to be proteins (protein)—chain compounds of about 20 natural amino acids (amino acid) RCH(NH2)COOH, ranging from the simplest, glycine, in which R is hydrogen, to tryptophan, in which R is

      Not only have methods been worked out for determining the amino acids found in an enzyme, but also the sequence of amino acids in an enzyme can be elucidated by a method developed by the English biochemist Frederick Sanger (Sanger, Frederick) in determining the structure of the protein hormone insulin. The first enzyme to have its complete amino acid sequence determined in this way was bovine pancreatic ribonuclease, which has 124 amino acids in its chain and a molecular weight of about 14,000; the enzyme catalyzes the degradation of ribonucleic acid, a substance active in protein synthesis in living cells. In January 1969 the synthesis of this same enzyme was reported from two different laboratories. The activity of an enzyme depends upon a three-dimensional, or tertiary, structure, but this, in turn, appears to depend solely upon the linear sequence of amino acids. The success of an enzyme's synthesis can be unequivocally checked by test of its enzymatic activity.

      Enzymes are extremely reactive, as can be shown with a very simple reaction—the splitting of hydrogen peroxide to form water and oxygen—brought about by colloidal metals (metal) and by the enzyme catalase. It has been found that one molecule of the latter will cause several million molecules of peroxide to decompose per minute, a rate comparable to that obtained with the best colloidal preparations. This speed of catalase decomposition is probably a maximum for enzymes. Slower-acting enzymes normally react at speeds of hundreds of reactions per minute. The rate of reaction is often expressed by an equation developed by L. Michaelis and M.L. Menten of the form

      in which V and K are constants for the particular enzymatic process, K being termed the Michaelis constant and [S] designated as the concentration of the reactant undergoing change. At low concentrations of S the rate is V[S]/K or proportional to the substrate concentration [S], whereas at high substrate concentrations the [S] terms cancel out and the reaction is essentially independent of the substrate concentration.

      A second characteristic of enzymes is their extreme specificity. It has been suggested that each biochemical process has its own specific enzyme. The biochemical processes induced by enzymes fall into broad classifications, such as hydrolysis, decomposition (or “splitting”), synthesis, and hydrogenation-dehydrogenation; as with catalysts in general, enzymes are active for both forward and reverse reactions.

      Like the laboratory catalysts, enzymes frequently have activators—coenzymes, which may be prosthetic groups (firmly bound to the enzyme itself), and inorganic ions (ion). adenosine triphosphate (ATP) is an important coenzyme participating in energy-producing processes and passage across cell membranes (membrane). Coenzymes often contain vitamins (vitamin) as part of their structure. calcium and magnesium ions are important enzyme activators. There are also many substances that inhibit, or poison, enzymes; cyanide ion is a potent inhibitor in many enzymic processes, as are nerve gases and insecticides (insecticide).

Sir Hugh S. Taylor

Additional Reading
Eric K. Rideal and Hugh S. Taylor, Catalysis in Theory and Practice, 2nd ed. (1926), recounts the state of this subject during World War I and the rapid advances in theory that occurred shortly thereafter. Paul H. Emmett, Catalysis Then and Now: A Survey of the Advances in Catalysis (1965), is a classic work written by a Nobel Prize winner. Paul H. Emmett (ed.), Catalysis, 7 vol. (1954–60), is a monumental treatment of the subject, both theory and practice, contributed by a group of international experts in the field. Additional information may be found in William P. Jencks, Catalysis in Chemistry and Enzymology (1969, reissued 1987); R. Pearce and W.R. Patterson (eds.), Catalysis and Chemical Processes (1981); John M. Thomas and Kirill I. Zamaraev (eds.), Perspectives in Catalysis (1992); Bruce C. Gates, Catalytic Chemistry (1992); and J.A. Moulijn, P.W.N.M. van Leeuwen, and R.A. van Santen (eds.), Catalysis: An Integrated Approach to Homogeneous, Heterogeneous, and Industrial Catalysis (1993). Advances in Catalysis (annual) records the year-to-year advances in the field as reported by international authorities.Sir Hugh S. Taylor

* * *


Universalium. 2010.

Игры ⚽ Поможем написать курсовую

Look at other dictionaries:

  • Catalysis — Ca*tal y*sis, n.; pl. {Catalyse}.[ML., fr. Gr. ? dissolution, fr. ? to destroy, dissolve; kata down, wholly + ? to loose.] [1913 Webster] 1. Dissolution; degeneration; decay. [R.] [1913 Webster] Sad catalysis and declension of piety. Evelyn.… …   The Collaborative International Dictionary of English

  • catalysis — (n.) 1650s, dissolution, from Gk. katalysis dissolution, a dissolving (of governments, military units, etc.), from katalyein to dissolve, from kata down (or completely ), see CATA (Cf. cata ), + lyein to loosen (see LOSE (Cf. lose)) …   Etymology dictionary

  • catalysis — ► NOUN ▪ the acceleration of a chemical reaction by a catalyst. DERIVATIVES catalytic adjective. ORIGIN Greek katalusis dissolution …   English terms dictionary

  • catalysis — [kə tal′ə sis] n. pl. catalyses [kə tal′əsēz΄] [ModL < Gr katalysis, dissolution: see CATA & LYSIS] the speeding up or, sometimes, slowing down of the rate of a chemical reaction caused by the addition of some substance that does not undergo a …   English World dictionary

  • Catalysis — Catalyst redirects here. For other uses, see Catalyst (disambiguation). Solid heterogeneous catalysts such as in automobile catalytic converters are plated on structures designed to maximize their surface area …   Wikipedia

  • catalysis — The effect that a catalyst exerts upon a chemical reaction. [G. katalysis, dissolution] contact c. a process wherein the catalyst is a solid and the catalyzed reaction is produced after the reactants (usually gases) have made contact with the… …   Medical dictionary

  • catalysis — [[t]kətæ̱lɪsɪs[/t]] N UNCOUNT Catalysis is the speeding up of a chemical reaction by adding a catalyst to it. [TECHNICAL] …   English dictionary

  • catalysis — katalizė statusas T sritis Standartizacija ir metrologija apibrėžtis Reakcijos greitinimas katalizatoriumi. atitikmenys: angl. catalysis vok. Katalyse, m rus. катализ, m pranc. catalyse, f …   Penkiakalbis aiškinamasis metrologijos terminų žodynas

  • catalysis — katalizė statusas T sritis chemija apibrėžtis Reakcijos greitinimas katalizatoriumi. atitikmenys: angl. catalysis rus. катализ …   Chemijos terminų aiškinamasis žodynas

  • catalysis — katalizė statusas T sritis fizika atitikmenys: angl. catalysis vok. Katalyse, f rus. катализ, m pranc. catalyse, f …   Fizikos terminų žodynas

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”