User:Redeemer079/Sandbox3

From Wikipedia, the free encyclopedia

Structures and mechanisms[edit]

Ribbon-diagram showing carbonic anhydrase II. The grey sphere is the zinc cofactor in the active site. Diagram drawn from PDB 1MOO.

Enzymes are generally globular proteins and range from just 62 amino acid residues in size, for the monomer of 4-oxalocrotonate tautomerase,[1] to over 2,500 residues in the animal fatty acid synthase.[2] A small number of RNA-based biological catalysts exist, with the most common being the ribosome, these are either referred to as RNA-enzymes, or ribozymes. The activities of enzymes are determined by their three-dimensional structure.[3] Most enzymes are much larger than the substrates they act on, and only a small portion of the enzyme (around 3–4 amino acids) is directly involved in catalysis.[4] 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 fold to produce a three-dimensional product. Each unique amino acid sequence produces a specific structure, which has unique properties. Individual protein chains may sometimes group together to form a protein complex. Most enzymes can be denatured—that is, unfolded and inactivated—by heating, which destroys the three-dimensional structure of the protein. Depending on the enzyme, denaturation may be reversible or irreversible.

Specificity[edit]

Enzymes are usually very specific as to which reactions they catalyze and the substrates that are involved in these reactions. Complementary shape, charge and hydrophilic/hydrophobic characteristics of enzymes and substrates are responsible for this specificity. Enzymes can also show impressive levels of stereospecificity, regioselectivity and chemoselectivity.[5]

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 DNA polymerase catalyzes a reaction in a first step and then checks that the product is correct in a second step.[6] This two-step process results in average error rates of less than 1 error in 100 million reactions in high-fidelity mammalian polymerases.[7] Similar proofreading mechanisms are also found in RNA polymerase,[8] aminoacyl tRNA synthetases[9] and ribosomes.[10]

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.[11]

"Lock and key" model[edit]

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.[12] 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[edit]

Diagrams to show the induced fit hypothesis of enzyme action.

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.[13] 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.[14] The active site continues to change until the substrate is completely bound, at which point the final shape and charge is determined.[15]

Mechanisms[edit]

Enzymes can act in several ways, all of which lower ΔG:[16]

  • 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,[17] and its contribution to catalysis is relatively small.[18]

Transition State Stabilization[edit]

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.[19] Such an environment does not exist in the uncatalyzed reaction in water.

Dynamics and function[edit]

Recent investigations have provided new insights into the connection between internal dynamics of enzymes and their mechanism of catalysis.[20][21][22] 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.[23][24][25][26] 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}).[27] 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[edit]

Allosteric enzymes change their structure in response to binding of effectors. Modulation can be direct, where the effector binds directly to binding sites in the enzyme, or indirect, where the effector binds to other proteins or protein subunits that interact with the allosteric enzyme and thus influence catalytic activity.

References[edit]

  1. ^ 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. doi:10.1016/S0021-9258(19)37101-7. PMID 1339435.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  2. ^ Smith S (1994). "The animal fatty acid synthase: one gene, one polypeptide, seven enzymes". FASEB J. 8 (15): 1248–59. doi:10.1096/fasebj.8.15.8001737. PMID 8001737. S2CID 22853095.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  3. ^ Anfinsen C.B. (1973). "Principles that Govern the Folding of Protein Chains". Science. 181 (4096): 223–230. doi:10.1126/science.181.4096.223. PMID 4124164.
  4. ^ The Catalytic Site Atlas at The European Bioinformatics Institute Accessed 04 April 2007
  5. ^ Jaeger KE, Eggert T. (2004). "Enantioselective biocatalysis optimized by directed evolution". Curr Opin Biotechnol. 15 (4): 305–313. doi:10.1016/j.copbio.2004.06.007. PMID 15358000.
  6. ^ Shevelev IV, Hubscher U. (2002). "The 3' 5' exonucleases". Nat Rev Mol Cell Biol. 3 (5): 364–376. doi:10.1038/nrm804. PMID 11988770. S2CID 31605786.
  7. ^ Berg J., Tymoczko J. and Stryer L. (2002) Biochemistry. W. H. Freeman and Company ISBN 0-7167-4955-6
  8. ^ Zenkin N, Yuzenkova Y, Severinov K. (2006). "Transcript-assisted transcriptional proofreading". Science. 313 (5786): 518–520. doi:10.1126/science.1127422. PMID 16873663. S2CID 40772789.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  9. ^ Ibba M, Soll D. (2000). "Aminoacyl-tRNA synthesis". Annu Rev Biochem. 69: 617–650. doi:10.1146/annurev.biochem.69.1.617. PMID 10966471.
  10. ^ Rodnina MV, Wintermeyer W. (2001). "Fidelity of aminoacyl-tRNA selection on the ribosome: kinetic and structural mechanisms". Annu Rev Biochem. 70: 415–435. doi:10.1146/annurev.biochem.70.1.415. PMID 11395413.
  11. ^ Firn, Richard. "The Screening Hypothesis - a new explanation of secondary product diversity and function". Retrieved 2006-10-11.
  12. ^ Fischer E. (1894). "Einfluss der Configuration auf die Wirkung der Enzyme". Ber. Dt. Chem. Ges. 27 (3): 2985–2993. doi:10.1002/cber.18940270364.
  13. ^ Koshland D. E. (1958). "Application of a Theory of Enzyme Specificity to Protein Synthesis". Proc. Natl. Acad. Sci. 44 (2): 98–104. doi:10.1073/pnas.44.2.98. PMC 335371. PMID 16590179.
  14. ^ Vasella A, Davies GJ, Bohm M. (2002). "Glycosidase mechanisms". Curr Opin Chem Biol. 6 (5): 619–629. doi:10.1016/S1367-5931(02)00380-0. PMID 12413546.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  15. ^ Boyer, Rodney (2002). "6". Concepts in Biochemistry (2nd ed.). New York, Chichester, Weinheim, Brisbane, Singapore, Toronto.: John Wiley & Sons, Inc. pp. 137–138. ISBN 0-470-00379-0.
  16. ^ Fersht, A (1985) Enzyme Structure and Mechanism (2nd ed) p50–52 W H Freeman & co, New York ISBN 0-7167-1615-1
  17. ^ Jencks W.P. "Catalysis in Chemistry and Enzymology." 1987, Dover, New York
  18. ^ 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. doi:10.1073/pnas.97.22.11899. PMC 17266. PMID 11050223.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  19. ^ Warshel A, Sharma PK, Kato M, Xiang Y, Liu H, Olsson MH (2006). "Electrostatic basis for enzyme catalysis". Chem. Rev. 106 (8): 3210–35. doi:10.1021/cr0503106. PMID 16895325.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  20. ^ Eisenmesser EZ, Bosco DA, Akke M, Kern D. Enzyme dynamics during catalysis. Science. 2002 February 22;295(5559):1520–3. PMID 11859194
  21. ^ 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
  22. ^ 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
  23. ^ Yang LW, Bahar I. (5 June 2005). "Coupling between catalytic site and collective dynamics: A requirement for mechanochemical activity of enzymes". Structure. 13 (6): 893–904. doi:10.1016/j.str.2005.03.015. PMC 1489920. PMID 15939021.{{cite journal}}: CS1 maint: date and year (link)
  24. ^ Agarwal PK, Billeter SR, Rajagopalan PT, Benkovic SJ, Hammes-Schiffer S. (5 March 2002). "Network of coupled promoting motions in enzyme catalysis". Proc. Natl. Acad. Sci. U S A. 99 (5): 2794–9. doi:10.1073/pnas.052005999. PMC 122427. PMID 11867722.{{cite journal}}: CS1 maint: date and year (link) CS1 maint: multiple names: authors list (link)
  25. ^ 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
  26. ^ Tousignant A, Pelletier JN. (Aug 2004). "Protein motions promote catalysis". Chem Biol. 11 (8): 1037–42. doi:10.1016/j.chembiol.2004.06.007. PMID 15324804.{{cite journal}}: CS1 maint: date and year (link)
  27. ^ 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