Enzyme



 

Enzymes are metabolic pathways occur in that cell.

Like all catalysts, enzymes work by lowering the artificial enzymes also display enzyme-like catalysis.[4]

Enzyme activity can be affected by other molecules. antibiotics. In addition, some household products use enzymes to speed up biochemical reactions (e.g., enzymes in biological washing powders break down protein or fat stains on clothes; enzymes in meat tenderizers break down proteins, making the meat easier to chew).

Etymology and history

  As early as the late 1700s and early 1800s, the digestion of meat by stomach secretions[5] and the conversion of starch to sugars by plant extracts and saliva were known. However, the mechanism by which this occurred had not been identified.[6]

In the 19th century, when studying the ferments", which were thought to function only within living organisms. He wrote that "alcoholic fermentation is an act correlated with the life and organization of the yeast cells, not with the death or putrefaction of the cells."[7]

In 1878 German physiologist Wilhelm Kühne (1837–1900) first used the term enzyme, which comes from Greek ενζυμον "in leaven", to describe this process. The word enzyme was used later to refer to nonliving substances such as pepsin, and the word ferment used to refer to chemical activity produced by living organisms.

In 1897 Eduard Buchner began to study the ability of yeast extracts that lacked any living yeast cells to ferment sugar. In a series of experiments at the University of Berlin, he found that the sugar was fermented even when there were no living yeast cells in the mixture.[8] He named the enzyme that brought about the fermentation of sucrose "DNA polymerase forms DNA polymers).

Having shown that enzymes could function outside a living cell, the next step was to determine their biochemical nature. Many early workers noted that enzymatic activity was associated with proteins, but several scientists (such as Nobel laureate Stanley, who worked on the digestive enzymes pepsin (1930), trypsin and chymotrypsin. These three scientists were awarded the 1946 Nobel Prize in Chemistry.[10]

This discovery that enzymes could be crystallized eventually allowed their structures to be solved by structural biology and the effort to understand how enzymes work at an atomic level of detail.

Structures and mechanisms

See also: Enzyme catalysis

 

Enzymes are generally globular proteins and range from just 62 amino acid residues in size, for the amino acids) is directly involved in catalysis.[15] The region that contains these catalytic residues, binds the substrate, and then carries out the reaction is known as the active site. Enzymes can also contain sites that bind cofactors, which are needed for catalysis. Some enzymes also have binding sites for small molecules, which are often direct or indirect products or substrates of the reaction catalyzed. This binding can serve to increase or decrease the enzyme's activity, providing a means for feedback regulation.

Like all proteins, enzymes are made as long, linear chains of amino acids that three-dimensional structure of the protein. Depending on the enzyme, denaturation may be reversible or irreversible.

Specificity

Enzymes are usually very specific as to which reactions they catalyze and the chemoselectivity.[16]

Some of the enzymes showing the highest specificity and accuracy are involved in the copying and expression of the genome. These enzymes have "proof-reading" mechanisms. Here, an enzyme such as ribosomes.[21]

Some enzymes that produce secondary metabolites are described as promiscuous, as they can act on a relatively broad range of different substrates. It has been suggested that this broad substrate specificity is important for the evolution of new biosynthetic pathways.[22]

"Lock and key" model

Enzymes are very specific, and it was suggested by Emil Fischer in 1894 that this was because both the enzyme and the substrate possess specific complementary geometric shapes that fit exactly into one another.[23] This is often referred to as "the lock and key" model. However, while this model explains enzyme specificity, it fails to explain the stabilization of the transition state that enzymes achieve. The "lock and key" model has proven inaccurate and the induced fit model is the most currently accepted enzyme-substrate-coenzyme figure.

Induced fit model

  In 1958 Daniel Koshland suggested a modification to the lock and key model: since enzymes are rather flexible structures, the active site is continually reshaped by interactions with the substrate as the substrate interacts with the enzyme.[24] As a result, the substrate does not simply bind to a rigid active site, the amino acid side chains which make up the active site are moulded into the precise positions that enable the enzyme to perform its catalytic function. In some cases, such as glycosidases, the substrate molecule also changes shape slightly as it enters the active site.[25] The active site continues to change until the substrate is completely bound, at which point the final shape and charge is determined.[26]

Mechanisms

Enzymes can act in several ways, all of which lower ΔG:[27]

  • Lowering the activation energy by creating an environment in which the transition state is stabilized (e.g. straining the shape of a substrate - by binding the transition-state conformation of the substrate/product molecules, the enzyme distorts the bound substrate(s) into their transition state form, thereby reducing the amount of energy required to complete the transition).
  • Lowering the energy of the transition state, but without distorting the substrate, by creating an environment with the opposite charge distribution to that of the transition state.
  • Providing an alternative pathway. For example,temporarily reacting with the substrate to form an intermediate ES complex, which would be impossible in the absence of the enzyme.
  • Reducing the reaction entropy change by bringing substrates together in the correct orientation to react. Considering ΔH alone overlooks this effect.

Interestingly, this entropic effect involves destabilization of the ground state,[28] and its contribution to catalysis is relatively small.[29]

Transition State Stabilization

The understanding of the origin of the reduction of ΔG requires one to find out how the enzymes can stabilize its transition state more than the transition state of the uncatalyzed reaction. Apparently, the most effective way for reaching large stabilization is the use of electrostatic effects, in particular, by having a relatively fixed polar environment that is oriented toward the charge distribution of the transition state.[30] Such an environment does not exist in the uncatalyzed reaction in water.

Dynamics and function

Recent investigations have provided new insights into the connection between internal dynamics of enzymes and their mechanism of catalysis.[31][32][33] An enzyme's internal dynamics are described as the movement of internal parts (e.g. amino acids, a group of amino acids, a loop region, an alpha helix, neighboring beta-sheets or even entire domain) of these biomolecules, which can occur at various time-scales ranging from femtoseconds to seconds. Networks of protein residues throughout an enzyme's structure can contribute to catalysis through dynamic motions.[34][35][36][37] Protein motions are vital to many enzymes, but whether small and fast vibrations or larger and slower conformational movements are more important depends on the type of reaction involved. These new insights also have implications in understanding allosteric effects and developing new drugs.

It should be clarified, however, that the dynamical time-dependent processes are not likely to help to accelerate enzymatic reactions, since such motions randomize and the rate constant is determined by the probability (P) of reaching the transition state, (P = exp {ΔG/RT}).[38] Furthermore, the reduction of ΔG requires having relatively smaller motions (in relation to the corresponding motions in solution reactions) for the transition between the reactant and the product states. Thus, it is not clear that motional or dynamical effects contribute to the catalysis of the chemical step.

Allosteric modulation

protein subunits that interact with the allosteric enzyme and thus influence catalytic activity.

Cofactors and coenzymes

Main articles: Cofactor (biochemistry) and Coenzyme

Cofactors

Some enzymes do not need any additional components to show full activity. However, others require non-protein molecules called cofactors to be bound for activity.[39] Cofactors can be either adenosine triphosphate. These molecules act to transfer chemical groups between enzymes.[40]

An example of an enzyme that contains a cofactor is redox reactions.

Enzymes that require a cofactor but do not have one bound are called apoenzymes. An apoenzyme together with its cofactor(s) is called a holoenzyme (this is the active form). Most cofactors are not covalently attached to an enzyme, but are very tightly bound. However, organic prosthetic groups can be covalently bound (e.g., thiamine pyrophosphate in the enzyme pyruvate dehydrogenase).

Coenzymes

  Coenzymes are small organic molecules that transport chemical groups from one enzyme to another.[42] Some of these chemicals such as S-adenosylmethionine.

Since coenzymes are chemically changed as a consequence of enzyme action, it is useful to consider coenzymes to be a special class of substrates, or second substrates, which are common to many different enzymes. For example, about 700 enzymes are known to use the coenzyme NADH.[43]

Coenzymes are usually regenerated and their concentrations maintained at a steady level inside the cell: for example, NADPH is regenerated through the pentose phosphate pathway and S-adenosylmethionine by methionine adenosyltransferase.

Thermodynamics

Main articles: Chemical equilibrium

 

As all catalysts, enzymes do not alter the position of the chemical equilibrium of the reaction. Usually, in the presence of an enzyme, the reaction runs in the same direction as it would without the enzyme, just more quickly. However, in the absence of the enzyme, other possible uncatalyzed, "spontaneous" reactions might lead to different products, because in those conditions this different product is formed faster.

Furthermore, enzymes can couple two or more reactions, so that a thermodynamically favorable reaction can be used to "drive" a thermodynamically unfavorable one. For example, the hydrolysis of ATP is often used to drive other chemical reactions.

Enzymes catalyze the forward and backward reactions equally. They do not alter the equilibrium itself, but only the speed at which it is reached. For example, carbonic anhydrase catalyzes its reaction in either direction depending on the concentration of its reactants.

\mathrm{CO_2 + H_2O {}^\mathrm{\quad Carbonic\ anhydrase} \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \overrightarrow{\qquad\qquad\qquad\qquad} H_2CO_3} (in tissues; high CO2 concentration)
\mathrm{H_2CO_3 {}^\mathrm{\quad Carbonic\ anhydrase} \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \overrightarrow{\qquad\qquad\qquad\qquad} CO_2 + H_2O} (in lungs; low CO2 concentration)

Nevertheless, if the equilibrium is greatly displaced in one direction, that is, in a very exergonic reaction, the reaction is effectively irreversible. Under these conditions the enzyme will, in fact, only catalyze the reaction in the thermodynamically allowed direction.

Kinetics

Main article: Enzyme kinetics

  Enzyme kinetics is the investigation of how enzymes bind substrates and turn them into products. The rate data used in kinetic analyses are obtained from enzyme assays.

In 1902 Victor Henri [44] proposed a quantitative theory of enzyme kinetics, but his experimental data were not useful because the significance of the hydrogen ion concentration was not yet appreciated. After Peter Lauritz Sørensen had defined the logarithmic pH-scale and introduced the concept of buffering in 1909[45] the German chemist Leonor Michaelis and his Canadian postdoc J. B. S. Haldane, who derived kinetic equations that are still widely used today.[47]

The major contribution of Henri was to think of enzyme reactions in two stages. In the first, the substrate binds reversibly to the enzyme, forming the enzyme-substrate complex. This is sometimes called the Michaelis complex. The enzyme then catalyzes the chemical step in the reaction and releases the product.

  Enzymes can catalyze up to several million reactions per second. For example, the reaction catalyzed by orotidine 5'-phosphate decarboxylase will consume half of its substrate in 78 million years if no enzyme is present. However, when the decarboxylase is added, the same process takes just 25 milliseconds.[48] Enzyme rates depend on solution conditions and substrate concentration. Conditions that denature the protein abolish enzyme activity, such as high temperatures, extremes of pH or high salt concentrations, while raising substrate concentration tends to increase activity. To find the maximum speed of an enzymatic reaction, the substrate concentration is increased until a constant rate of product formation is seen. This is shown in the saturation curve on the right. Saturation happens because, as substrate concentration increases, more and more of the free enzyme is converted into the substrate-bound ES form. At the maximum velocity (Vmax) of the enzyme, all the enzyme active sites are bound to substrate, and the amount of ES complex is the same as the total amount of enzyme. However, Vmax is only one kinetic constant of enzymes. The amount of substrate needed to achieve a given rate of reaction is also important. This is given by the Michaelis-Menten constant (Km), which is the substrate concentration required for an enzyme to reach one-half its maximum velocity. Each enzyme has a characteristic Km for a given substrate, and this can show how tight the binding of the substrate is to the enzyme. Another useful constant is kcat, which is the number of substrate molecules handled by one active site per second.

The efficiency of an enzyme can be expressed in terms of kcat/Km. This is also called the specificity constant and incorporates the superoxide dismutase.

Michaelis-Menten kinetics relies on the Michaelis-Menten kinetics may be applied.[50][51][52][53]

Some enzymes operate with kinetics which are faster than diffusion rates, which would seem to be impossible. Several mechanisms have been invoked to explain this phenomenon. Some proteins are believed to accelerate catalysis by drawing their substrate in and pre-orienting them by using dipolar electric fields. Other models invoke a quantum-mechanical tunneling explanation, whereby a proton or an electron can tunnel through activation barriers, although for proton tunneling this model remains somewhat controversial.[54][55] Quantum tunneling for protons has been observed in tryptamine.[56] This suggests that enzyme catalysis may be more accurately characterized as "through the barrier" rather than the traditional model, which requires substrates to go "over" a lowered energy barrier.

Inhibition

 

 

Main article: Enzyme inhibitor

Enzyme reaction rates can be decreased by various types of enzyme inhibitors.

Competitive inhibition

In competitive inhibition, the inhibitor and substrate compete for the enzyme (i.e., they can not bind at the same time). Often competitive inhibitors strongly resemble the real substrate of the enzyme. For example, tetrahydrofolate. The similarity between the structures of folic acid and this drug are shown in the figure to the right bottom. Note that binding of the inhibitor need not be to the substrate binding site (as frequently stated), if binding of the inhibitor changes the conformation of the enzyme to prevent substrate binding and vice versa. In competitive inhibition the maximal velocity of the reaction is not changed, but higher substrate concentrations are required to reach a given velocity, increasing the apparent Km.

Uncompetitive inhibition

In uncompetitive inhibition the inhibitor can not bind to the free enzyme, but only to the ES-complex. The EIS-complex thus formed is enzymatically inactive. This type of inhibition is rare, but may occur in multimeric enzymes.

Non-competitive inhibition

Non-competitive inhibitors can bind to the enzyme at the same time as the substrate, i.e. they never bind to the active site. Both the EI and EIS complexes are enzymatically inactive. Because the inhibitor can not be driven from the enzyme by higher substrate concentration (in contrast to competitive inhibition), the apparent Vmax changes. But because the substrate can still bind to the enzyme, the Km stays the same.

Mixed inhibition

This type of inhibition resembles the non-competitive, except that the EIS-complex has residual enzymatic activity.

In many organisms inhibitors may act as part of a feedback mechanism. If an enzyme produces too much of one substance in the organism, that substance may act as an inhibitor for the enzyme at the beginning of the pathway that produces it, causing production of the substance to slow down or stop when there is sufficient amount. This is a form of negative feedback. Enzymes which are subject to this form of regulation are often multimeric and have allosteric binding sites for regulatory substances. Their substrate/velocity plots are not hyperbolar, but sigmoidal (S-shaped).

 

Aspirin also act in this manner. With these drugs, the compound is bound in the active site and the enzyme then converts the inhibitor into an activated form that reacts irreversibly with one or more amino acid residues.

Uses of inactivators

Inhibitors are often used as drugs, but they can also act as poisons. However, the difference between a drug and a poison is usually only a matter of amount, since most drugs are toxic at some level, as antibiotics and other anti-infective drugs are just specific poisons that kill a pathogen but not its host.

An example of an inactivator being used as a drug is cellular respiration.[60]

Biological function

Enzymes serve a wide variety of functions inside living organisms. They are indispensable for neuraminidase.

An important function of enzymes is in the digestive systems of animals. Enzymes such as cellulase to break down the cellulose cell walls of plant fiber.[64]

Several enzymes can work together in a specific order, creating metabolic pathways. In a metabolic pathway, one enzyme takes the product of another enzyme as a substrate. After the catalytic reaction, the product is then passed on to another enzyme. Sometimes more than one enzyme can catalyse the same reaction in parallel, this can allow more complex regulation: with for example a low constant activity being provided by one enzyme but an inducible high activity from a second enzyme.

Enzymes determine what steps occur in these pathways. Without enzymes, metabolism would neither progress through the same steps, nor be fast enough to serve the needs of the cell. Indeed, a metabolic pathway such as glucose-6-phosphate is found to be the only significant product. Consequently, the network of metabolic pathways within each cell depends on the set of functional enzymes that are present.

Control of activity

There are five main ways that enzyme activity is controlled in the cell.

  1. Enzyme production (drug interactions.
  2. Enzymes can be compartmentalized, with different metabolic pathways occurring in different cellular compartments. For example, fatty acids are synthesized by one set of enzymes in the cytosol, endoplasmic reticulum and the Golgi apparatus and used by a different set of enzymes as a source of energy in the β-oxidation.[65]
  3. Enzymes can be regulated by inhibitors and activators. For example, the end product(s) of a metabolic pathway are often inhibitors for one of the first enzymes of the pathway (usually the first irreversible step, called committed step), thus regulating the amount of end product made by the pathways. Such a regulatory mechanism is called a negative feedback mechanism, because the amount of the end product produced is regulated by its own concentration. Negative feedback mechanism can effectively adjust the rate of synthesis of intermediate metabolites according to the demands of the cells. This helps allocate materials and energy economically, and prevents the manufacture of excess end products. Like other homeostatic devices, the control of enzymatic action helps to maintain a stable internal environment in living organisms.
  4. Enzymes can be regulated through pancreas and transported in this form to the stomach where it is activated. This stops the enzyme from digesting the pancreas or other tissues before it enters the gut. This type of inactive precursor to an enzyme is known as a zymogen.
  5. Some enzymes may become activated when localized to a different environment (eg. from a reducing (cytoplasm) to an oxidising (periplasm) environment, high pH to low pH etc). For example, hemagglutinin of the influenza virus undergoes a conformational change once it encounters the acidic environment of the host cell vesicle causing its activation.[67]

Involvement in disease

  Since the tight control of enzyme activity is essential for homeostasis, any malfunction (mutation, overproduction, underproduction or deletion) of a single critical enzyme can lead to a genetic disease. The importance of enzymes is shown by the fact that a lethal illness can be caused by the malfunction of just one type of enzyme out of the thousands of types present in our bodies.

One example is the most common type of phenylalanine, results in build-up of phenylalanine and related products. This can lead to mental retardation if the disease is untreated.[68]

Another example is when germline mutations in genes coding for DNA repair enzymes cause hereditary cancer syndromes such as xeroderma pigmentosum. Defects in these enzymes cause cancer since the body is less able to repair mutations in the genome. This causes a slow accumulation of mutations and results in the development of many types of cancer in the sufferer.

Naming conventions

An enzyme's name is often derived from its substrate or the chemical reaction it catalyzes, with the word ending in -ase. Examples are glucose into the sweetener fructose, is a xylose isomerase in vivo.

The International Union of Biochemistry and Molecular Biology have developed a nomenclature for enzymes, the EC numbers; each enzyme is described by a sequence of four numbers preceded by "EC". The first number broadly classifies the enzyme based on its mechanism:

The top-level classification is

The complete nomenclature can be browsed at http://www.chem.qmul.ac.uk/iubmb/enzyme/.

Industrial applications

Enzymes are used in the protein engineering is an active area of research and involves attempts to create new enzymes with novel properties, either through rational design or in vitro evolution.[69][70]

Application Enzymes used Uses
Baking industry
 
Fungal alpha-amylase enzymes are normally inactivated at about 50 degrees Celsius, but are destroyed during the baking process. Catalyze breakdown of starch in the flour to sugar. Yeast action on sugar produces carbon dioxide. Used in production of white bread, buns, and rolls.
Proteases Biscuit manufacturers use them to lower the protein level of flour.
Baby foods Trypsin To predigest baby foods.
Brewing industry
 
Enzymes from barley are released during the mashing stage of beer production. They degrade starch and proteins to produce simple sugar, amino acids and peptides that are used by yeast for fermentation.
Industrially produced barley enzymes Widely used in the brewing process to substitute for the natural enzymes found in barley.
Amylase, glucanases, proteases Split polysaccharides and proteins in the malt.
Betaglucanases and arabinoxylanases Improve the wort and beer filtration characteristics.
Amyloglucosidase and pullulanases Low-calorie beer and adjustment of fermentability.
Proteases Remove cloudiness produced during storage of beers.
Acetolactatedecarboxylase (ALDC) Avoid the formation of diacetyl
Fruit juices Cellulases, pectinases Clarify fruit juices
Dairy industry
 
Rennin, derived from the stomachs of young ruminant animals (like calves and lambs). Manufacture of cheese, used to hydrolyze protein.
Microbially produced enzyme Now finding increasing use in the dairy industry.
Lipases Is implemented during the production of Roquefort cheese to enhance the ripening of the blue-mould cheese.
Lactases Break down glucose and galactose.
Meat tenderizers Papain To soften meat for cooking.
Starch industry
Glucose
Fructose
Amylases, amyloglucosideases and glucoamylases Converts glucose and various syrups.
Glucose isomerase Converts glucose into fructose in production of high fructose syrups from starchy materials. These syrups have enhanced sweetening properties and lower calorific values than sucrose for the same level of sweetness.
Paper industry
 
Cellulases and ligninases Degrade starch to lower viscosity, aiding sizing and coating paper. Xylanases reduce bleach required for decolorising; cellulases smooth fibers, enhance water drainage, and promote ink removal; lipases reduce pitch and lignin-degrading enzymes remove lignin to soften paper.
Biofuel industry
 
Cellulases Used to break down cellulose into sugars that can be fermented (see cellulosic ethanol).
Ligninases Use of lignin waste
Biological detergent
 
Primarily proteases, produced in an extracellular form from bacteria Used for presoak conditions and direct liquid applications helping with removal of protein stains from clothes.
Amylases Detergents for machine dish washing to remove resistant starch residues.
Lipases Used to assist in the removal of fatty and oily stains.
Cellulases Used in biological fabric conditioners.
Contact lens cleaners Proteases To remove proteins on contact lens to prevent infections.
Rubber industry Catalase To generate peroxide to convert latex into foam rubber.
Photographic industry Protease (ficin) Dissolve silver content.
Molecular biology
 
Restriction enzymes, DNA ligase and polymerases Used to manipulate DNA in genetic engineering, important in pharmacology, agriculture and medicine. Essential for restriction digestion and the polymerase chain reaction. Molecular biology is also important in forensic science.

See also

References

  1. ^ Smith AD (Ed) et al. (1997) Oxford Dictionary of Biochemistry and Molecular Biology Oxford University Press ISBN 0-19-854768-4
  2. ^ Bairoch A. (2000). "The ENZYME database in 2000". Nucleic Acids Res 28: 304–305. PMID 10592255.
  3. ^ Lilley D (2005). "Structure, folding and mechanisms of ribozymes". Curr Opin Struct Biol 15 (3): 313-23. PMID 15919196.
  4. ^ Groves JT (1997). "Artificial enzymes. The importance of being selective". Nature 389 (6649): 329-30. PMID 9311771.
  5. ^ de Réaumur, RAF (1752). "Observations sur la digestion des oiseaux". Histoire de l'academie royale des sciences 1752: 266, 461.
  6. ^ Williams, H. S. (1904) A History of Science: in Five Volumes. Volume IV: Modern Development of the Chemical and Biological Sciences Harper and Brothers (New York) Accessed 04 April 2007
  7. ^ Dubos J. (1951). "Louis Pasteur: Free Lance of Science, Gollancz. Quoted in Manchester K. L. (1995) Louis Pasteur (1822–1895)—chance and the prepared mind.". Trends Biotechnol 13 (12): 511–515. PMID 8595136.
  8. ^ Nobel Laureate Biography of Eduard Buchner at http://nobelprize.org Accessed 04 April 2007
  9. ^ Text of Eduard Buchner's 1907 Nobel lecture at http://nobelprize.org Accessed 04 April 2007
  10. ^ 1946 Nobel prize for Chemistry laureates at http://nobelprize.org Accessed 04 April 2007
  11. ^ Blake CC, Koenig DF, Mair GA, North AC, Phillips DC, Sarma VR. (1965). "Structure of hen egg-white lysozyme. A three-dimensional Fourier synthesis at 2 Angstrom resolution.". Nature 22 (206): 757–761. PMID 5891407.
  12. ^ Chen LH, Kenyon GL, Curtin F, Harayama S, Bembenek ME, Hajipour G, Whitman CP (1992). "4-Oxalocrotonate tautomerase, an enzyme composed of 62 amino acid residues per monomer". J. Biol. Chem. 267 (25): 17716-21. PMID 1339435.
  13. ^ Smith S (1994). "The animal fatty acid synthase: one gene, one polypeptide, seven enzymes". FASEB J. 8 (15): 1248–59. PMID 8001737.
  14. ^ Anfinsen C.B. (1973). "Principles that Govern the Folding of Protein Chains". Science: 223–230. PMID 4124164.
  15. ^ The Catalytic Site Atlas at The European Bioinformatics Institute Accessed 04 April 2007
  16. ^ Jaeger KE, Eggert T. (2004). "Enantioselective biocatalysis optimized by directed evolution.". Curr Opin Biotechnol. 15(4): 305–313. PMID 15358000.
  17. ^ Shevelev IV, Hubscher U. (2002). "The 3' 5' exonucleases.". Nat Rev Mol Cell Biol. 3 (5): 364–376. PMID 11988770.
  18. ^ Berg J., Tymoczko J. and Stryer L. (2002) Biochemistry. W. H. Freeman and Company ISBN 0-7167-4955-6
  19. ^ Zenkin N, Yuzenkova Y, Severinov K. (2006). "Transcript-assisted transcriptional proofreading.". Science. 313: 518–520. PMID 16873663.
  20. ^ Ibba M, Soll D. (2000). "Aminoacyl-tRNA synthesis.". Annu Rev Biochem. 69: 617–650. PMID 10966471.
  21. ^ Rodnina MV, Wintermeyer W. (2001). "Fidelity of aminoacyl-tRNA selection on the ribosome: kinetic and structural mechanisms.". Annu Rev Biochem. 70: 415–435. PMID 11395413.
  22. ^ Firn, Richard. The Screening Hypothesis - a new explanation of secondary product diversity and function. Retrieved on 2006-10-11.
  23. ^ Fischer E. (1894). "Einfluss der Configuration auf die Wirkung der Enzyme". Ber. Dt. Chem. Ges. 27: 2985–2993.
  24. ^ Koshland D. E. (1958). "Application of a Theory of Enzyme Specificity to Protein Synthesis". Proc. Natl. Acad. Sci. 44 (2): 98–104. PMID 16590179.
  25. ^ Vasella A, Davies GJ, Bohm M. (2002). "Glycosidase mechanisms.". Curr Opin Chem Biol. 6 (5): 619–629. PMID 12413546.
  26. ^ Boyer, Rodney [2002]. "6", Concepts in Biochemistry, 2nd ed. (in English), New York, Chichester, Weinheim, Brisbane, Singapore, Toronto.: John Wiley & Sons, Inc., 137–138. ISBN 0-470-00379-0. 
  27. ^ Fersht, A (1985) Enzyme Structure and Mechanism (2nd ed) p50–52 W H Freeman & co, New York ISBN 0-7167-1615-1
  28. ^ Jencks W.P. "Catalysis in Chemistry and Enzymology." 1987, Dover, New York
  29. ^ Villa J, Strajbl M, Glennon TM, Sham YY, Chu ZT, Warshel A (2000). "How important are entropic contributions to enzyme catalysis?". Proc. Natl. Acad. Sci. U.S.A. 97 (22): 11899-904. PMID 11050223.
  30. ^ Warshel A, Sharma PK, Kato M, Xiang Y, Liu H, Olsson MH (2006). "Electrostatic basis for enzyme catalysis". Chem. Rev. 106 (8): 3210-35. PMID 16895325.
  31. ^ Eisenmesser EZ, Bosco DA, Akke M, Kern D. Enzyme dynamics during catalysis. Science. 2002 February 22;295(5559):1520–3. PMID: 11859194
  32. ^ Agarwal PK. Role of protein dynamics in reaction rate enhancement by enzymes. J Am Chem Soc. 2005 November 2;127(43):15248-56. PMID: 16248667
  33. ^ Eisenmesser EZ, Millet O, Labeikovsky W, Korzhnev DM, Wolf-Watz M, Bosco DA, Skalicky JJ, Kay LE, Kern D. Intrinsic dynamics of an enzyme underlies catalysis. Nature. 2005 November 3;438(7064):117-21. PMID: 16267559
  34. ^ Yang LW, Bahar I. (June 2005). "Coupling between catalytic site and collective dynamics: A requirement for mechanochemical activity of enzymes.". Structure. 13: 893–904. PMID 15939021.
  35. ^ Agarwal PK, Billeter SR, Rajagopalan PT, Benkovic SJ, Hammes-Schiffer S. (March 2002). "Network of coupled promoting motions in enzyme catalysis.". Proc. Natl. Acad. Sci. U S A. 99: 2794–9. PMID 11867722.
  36. ^ Agarwal PK, Geist A, Gorin A. Protein dynamics and enzymatic catalysis: investigating the peptidyl-prolyl cis-trans isomerization activity of cyclophilin A. Biochemistry. 2004 August 24;43(33):10605-18. PMID: 15311922
  37. ^ Tousignant A, Pelletier JN. (Aug 2004). "Protein motions promote catalysis.". Chem Biol. 11 (8): 1037–42. PMID 15324804.
  38. ^ Olsson M.H.M., Parson W.W. and Warshel A. "Dynamical Contributions to Enzyme Catalysis: Critical Tests of A Popular Hypothesis" Chem. Rev., 2006 105: 1737-1756
  39. ^ de Bolster, M.W.G. (1997). Glossary of Terms Used in Bioinorganic Chemistry. International Union of Pure and Applied Chemistry. Retrieved on 2007-10-30.
  40. ^ de Bolster, M.W.G. (1997). Glossary of Terms Used in Bioinorganic Chemistry. International Union of Pure and Applied Chemistry. Retrieved on 2007-10-30.
  41. ^ Fisher Z, Hernandez Prada JA, Tu C, Duda D, Yoshioka C, An H, Govindasamy L, Silverman DN and McKenna R. (2005). "Structural and kinetic characterization of active-site histidine as a proton shuttle in catalysis by human carbonic anhydrase II.". Biochemistry. 44(4): 1097-115. PMID 15667203.
  42. ^ AF Wagner, KA Folkers (1975) Vitamins and coenzymes. Interscience Publishers New York| ISBN 0-88275-258-8
  43. ^ BRENDA The Comprehensive Enzyme Information System Accessed 04 April 2007
  44. ^ Henri, V. (1902). "Theorie generale de l'action de quelques diastases". Compt. rend. hebd. Acad. Sci. Paris 135: 916-919.
  45. ^ Sørensen,P.L. (1909). "Enzymstudien {II}. Über die Messung und Bedeutung der Wasserstoffionenkonzentration bei enzymatischen Prozessen". Biochem. Z. 21: 131-304.
  46. ^ Michaelis L., Menten M. (1913). "Die Kinetik der Invertinwirkung". Biochem. Z. 49: 333–369. English translation Accessed 6 April 2007
  47. ^ Briggs G. E., Haldane J. B. S. (1925). "A note on the kinetics of enzyme action". Biochem. J. 19: 339–339. PMID 16743508.
  48. ^ Radzicka A, Wolfenden R. (1995). "A proficient enzyme.". Science 6 (267): 90–931. PMID 7809611.
  49. ^ Ellis RJ (2001). "Macromolecular crowding: obvious but underappreciated". Trends Biochem. Sci. 26 (10): 597-604. PMID 11590012.
  50. ^ Kopelman R (1988). "Fractal Reaction Kinetics". Science 241 (4873): 1620–26.
  51. ^ Savageau MA (1995). "Michaelis-Menten mechanism reconsidered: implications of fractal kinetics". J. Theor. Biol. 176 (1): 115-24. PMID 7475096.
  52. ^ Schnell S, Turner TE (2004). "Reaction kinetics in intracellular environments with macromolecular crowding: simulations and rate laws". Prog. Biophys. Mol. Biol. 85 (2–3): 235-60. PMID 15142746.
  53. ^ Xu F, Ding H (2007). "A new kinetic model for heterogeneous (or spatially confined) enzymatic catalysis: Contributions from the fractal and jamming (overcrowding) effects". Appl. Catal. A: Gen. 317 (1): 70–81. doi:10.1016/j.apcata.2006.10.014.
  54. ^ Garcia-Viloca M., Gao J., Karplus M., Truhlar D. G. (2004). "How enzymes work: analysis by modern rate theory and computer simulations.". Science 303 (5655): 186–195. PMID 14716003.
  55. ^ Olsson M. H., Siegbahn P. E., Warshel A. (2004). "Simulations of the large kinetic isotope effect and the temperature dependence of the hydrogen atom transfer in lipoxygenase". J. Am. Chem. Soc. 126 (9): 2820-1828. PMID 14995199.
  56. ^ Masgrau L., Roujeinikova A., Johannissen L. O., Hothi P., Basran J., Ranaghan K. E., Mulholland A. J., Sutcliffe M. J., Scrutton N. S., Leys D. (2006). "Atomic Description of an Enzyme Reaction Dominated by Proton Tunneling". Science 312 (5771): 237–241. PMID 16614214.
  57. ^ Cleland, W.W. (1963). "The Kinetics of Enzyme-catalyzed Reactions with two or more Substrates or Products 2. {I}nhibition: Nomenclature and Theory". Biochim. Biophys. Acta 67: 173-187.
  58. ^ Poulin R, Lu L, Ackermann B, Bey P, Pegg AE. Mechanism of the irreversible inactivation of mouse ornithine decarboxylase by alpha-difluoromethylornithine. Characterization of sequences at the inhibitor and coenzyme binding sites. J Biol Chem. 1992 Jan 5;267(1):150–8. PMID 1730582
  59. ^ Ball, Philip (2006) The Devil's Doctor: Paracelsus and the World of Renaissance Magic and Science. Farrar, Straus and Giroux ISBN 0-374-22979-1
  60. ^ Yoshikawa S and Caughey WS. (May 1990). "Infrared evidence of cyanide binding to iron and copper sites in bovine heart cytochrome c oxidase. Implications regarding oxygen reduction.". J Biol Chem. 265 (14): 7945–7958. PMID 2159465.
  61. ^ Hunter T. (1995). "Protein kinases and phosphatases: the yin and yang of protein phosphorylation and signaling.". Cell. 80(2): 225–236. PMID 7834742.
  62. ^ Berg JS, Powell BC, Cheney RE. (2001). "A millennial myosin census.". Mol Biol Cell. 12(4): 780–794. PMID 11294886.
  63. ^ Meighen EA. (1991). "Molecular biology of bacterial bioluminescence.". Microbiol Rev. 55(1): 123–142. PMID 2030669.
  64. ^ Mackie RI, White BA (1990). "Recent advances in rumen microbial ecology and metabolism: potential impact on nutrient output". J. Dairy Sci. 73 (10): 2971–95. PMID 2178174.
  65. ^ Faergeman N. J, Knudsen J. (April 1997). "Role of long-chain fatty acyl-CoA esters in the regulation of metabolism and in cell signalling". Biochem J 323: 1–12. PMID 9173866.
  66. ^ Doble B. W., Woodgett J. R. (April 2003). "GSK-3: tricks of the trade for a multi-tasking kinase". J. Cell. Sci. 116: 1175–1186. PMID 12615961.
  67. ^ Carr C. M., Kim P. S. (April 2003). "A spring-loaded mechanism for the conformational change of influenza hemagglutinin". Cell 73: 823–832. PMID 8500173.
  68. ^ Phenylketonuria: NCBI Genes and Disease Accessed 04 April 2007
  69. ^ Renugopalakrishnan V, Garduno-Juarez R, Narasimhan G, Verma CS, Wei X, Li P. (2005). "Rational design of thermally stable proteins: relevance to bionanotechnology.". J Nanosci Nanotechnol. 5 (11): 1759–1767. PMID 16433409.
  70. ^ Hult K, Berglund P. (2003). "Engineered enzymes for improved organic synthesis.". Curr Opin Biotechnol. 14 (4): 395–400. PMID 12943848.

Further reading

Etymology and history

  • New Beer in an Old Bottle: Eduard Buchner and the Growth of Biochemical Knowledge, edited by Athel Cornish-Bowden and published by Universitat de València (1997): ISBN 84-370-3328-4, A history of early enzymology.
  • Williams, Henry Smith, 1863–1943. A History of Science: in Five Volumes. Volume IV: Modern Development of the Chemical and Biological Sciences, A textbook from the 19th century.
  • Kleyn, J. and Hough J. The Microbiology of Brewing. Annual Review of Microbiology (1971) Vol. 25: 583–608

Enzyme structure and mechanism

  • Fersht, A. Structure and Mechanism in Protein Science: A Guide to Enzyme Catalysis and Protein Folding. W. H. Freeman, 1998 ISBN 0-7167-3268-8
  • Walsh, C., Enzymatic Reaction Mechanisms. W. H. Freeman and Company. 1979. ISBN 0-7167-0070-0
  • Page, M. I., and Williams, A. (Eds.), 1987. Enzyme Mechanisms. Royal Society of Chemistry. ISBN 0-85186-947-5
  • Bugg, T. Introduction to Enzyme and Coenzyme Chemistry, 2004, Blackwell Publishing Limited; 2nd edition. ISBN 1-40511-452-5
  • Warshel, A., Computer Modeling of Chemical Reactions in enzymes and Solutions John Wiley & Sons Inc. 1991. ISBN 0-471-18440-3

Thermodynamics

  • Reactions and Enzymes Chapter 10 of On-Line Biology Book at Estrella Mountain Community College.

Kinetics and inhibition

  • Athel Cornish-Bowden, Fundamentals of Enzyme Kinetics. (3rd edition), Portland Press (2004), ISBN 1-85578-158-1.
  • Irwin H. Segel, Enzyme Kinetics: Behavior and Analysis of Rapid Equilibrium and Steady-State Enzyme Systems. Wiley-Interscience; New Ed edition (1993), ISBN 0-471-30309-7.
  • John W. Baynes, Medical Biochemistry, Elsevier-Mosby; 2th Edition (2005), ISBN 0-7234-3341-0, p. 57.

Function and control of enzymes in the cell

  • Price, N. and Stevens, L., Fundamentals of Enzymology: Cell and Molecular Biology of Catalytic Proteins Oxford University Press, (1999), ISBN 0-19-850229-X
  • Nutritional and Metabolic Diseases Chapter of the on-line textbook "Introduction to Genes and Disease" from the NCBI.

Enzyme-naming conventions

  • Enzyme Nomenclature, Recommendations for enzyme names from the Nomenclature Committee of the International Union of Biochemistry and Molecular Biology.
  • Koshland D. The Enzymes, v. I, ch. 7, Acad. Press, New York, (1959)

Industrial applications

  • History of industrial enzymes, Article about the history of industrial enzymes, from the late 1900s to the present times.

 
This article is licensed under the GNU Free Documentation License. It uses material from the Wikipedia article "Enzyme". A list of authors is available in Wikipedia.