Home About Us Contact Us

 

Table of Content - Volume 13 Issue 1 - January 2020




RNA as ENZYMES and comparison of its properties with PROTEINS as ENZYMES

 

Kavindra Borgaonkar1*, Ranjit Patil2

 

1Associate Professor, Department of Biochemistry, Government Medical College Latur, Maharashtra, INDIA.

2Professor, Department of Medical Biochemistry, Ssr Medical College Belle Rive Mauritius.

Email: kb2172@rediffmail.com, ranj42@gmail.com

 

Abstract               Background: Enzymes /ˈɛnzaɪmz/ are both proteins and biological catalysts (biocatalysts). Catalysts accelerate chemical reactions. The molecules upon which enzymes may act are called substrates, and the enzyme converts the substrates into different molecules known as products. Almost all metabolic processes in the cell need enzyme catalysis in order to occur at rates fast enough to sustain life. RNA- The chemically versatile molecule : of The information content of the cell is contained in DNA, which codes for proteins that carry out the majority of cellular functions. However, the discovery of catalytic RNA led to the RNA world hypothesis: at one point, RNA could have been both the information carrier and the functional molecule . It is necessary but not sufficient for the RNA world to be self-perpetuating: it must also have given rise to current biology featuring protein enzymes. The details of this transition are purely speculative, but it may have included the recruitment of amino acids or short peptides as enzymatic cofactors. Clearly, RNA is capable of specifically binding these ligands and could use them to assist in chemistry. Enzymes are generally globular proteins, acting alone or in larger complexes. The sequence of the amino acids specifies the structure which in turn determines the catalytic activity of the enzyme . Dynamics of Protein –Enzymes : These are not rigid, static structures; instead they have complex internal dynamic motions – that is, movements of parts of the enzyme's structure such as individual amino acid residues, groups of residues forming a protein loop or unit of secondary structure, or even an entire protein domain. Types of inhibition-Competitive, Non-competitive, Uncompetitive, Mixed, Irreversible. Conclusion: Only two reactions are observed to be catalysed by RNA in nature. However, the diversity of catalytic mechanisms for these reactions implies that RNA may be capable of much more. This is supported by the success of in vitro selection to develop ribozymes for RNA ligation.

Key Word: RNA – Enzymes , Protein Enzymes , ribozymes, properties of enzymes

 

 

INTRODUCTION

Enzymes /ˈɛnzaɪmz/ are both proteins and biological catalysts (biocatalysts). Catalysts accelerate chemical reactions. The molecules upon which enzymes may act are called substrates, and the enzyme converts the substrates into different molecules known as products. Almost all metabolic processes in the cell need enzyme catalysis in order to occur at rates fast enough to sustain life.1:8.1 Metabolic pathways depend upon enzymes to catalyze individual steps. The study of enzymes is called enzymology and a new field of pseudoenzyme analysis has recently grown up, recognising that during evolution, some enzymes have lost the ability to carry out biological catalysis, which is often reflected in their amino acid sequences and unusual 'pseudocatalytic' properties.2,3 Enzymes are known to catalyze more than 5,000 biochemical reaction types.4 Other biocatalysts are catalytic RNA molecules, called ribozymes. Enzymes' specificity comes from their unique three-dimensional structures. Like all catalysts, enzymes increase the reaction rate by lowering its activation energy. Some enzymes can make their conversion of substrate to product occur many millions of times faster. An extreme example is orotidine 5'-phosphate decarboxylase, which allows a reaction that would otherwise take millions of years to occur in milliseconds.5,6 RNA- The chemically versatile molecule : of The information content of the cell is contained in DNA, which codes for proteins that carry out the majority of cellular functions. However, the discovery of catalytic RNA led to the RNA world hypothesis: at one point, RNA could have been both the information carrier and the functional molecule 7. At first the main function would be self-replication, but as more complex RNAs evolved, they would need to synthesize their precursors, store and use energy, and isolate their reaction products from the environment. Eventually, RNA would have to transition towards modern biology, where proteins play most of the catalytic roles. This would involve the increasingly complex task of message-directed protein synthesis, including the specific activation of amino acids to be incorporated. It is now known that natural ribozymes use a variety of catalytic mechanisms, even for phosphoryl transfer. Furthermore, RNA regulatory elements, called riboswitches, have been shown to bind a variety of cofactors, small molecules which expand the repertoire of modern protein enzymes. In addition to phosphoryl transfer ribozymes, the RNA component of the ribosome is responsible for catalysing protein synthesis. This is the key reaction required to transition from RNA to protein as functional molecule. These observations indicate the versatility of RNA in catalysing chemical reactions, an essential characteristic for the RNA world. The RNA components of other phosphoryl transfer enzymes—RNase P and the spliceosome—also bind divalent metal ions important for catalysis 8,9. Clearly, RNA is suited for recruiting divalent metal ions, which can be used for a variety of catalytic roles. The large ribozymes are all metalloenzymes, but surprisingly, the hairpin, Varkud Satellite (VS) and hammerhead ribozymes are active in the absence of divalent metal ions 10. Each of these ribozymes, along with the hepatitis delta virus ribozyme (HDV), catalyse self-cleavage and ligation reactions as part of processing genome replicants. The hammerhead ribozyme is perhaps the most extensively studied small ribozyme, and the wealth of biochemical and structural data support a catalytic mechanism dependent on RNA functional groups. Although early crystal structures of a minimal hammerhead construct conflicted with biochemical data, this has been reconciled by the structure of a full-length, fully active construct.11 This appears to be a common strategy for the small ribozymes. These experiments implicate nucleobases in acid–base catalysis for the hairpin, HDV and VS ribozymes 13,14. This is unexpected because no RNA functional groups have unperturbed pKas near neutrality. It is possible that some of these bases are participating in hydrogen bonds instead of full acid–base catalysis. Alternatively, the bases could be functioning as alternate tautomers, especially for guanine, which has been repeatedly observed in position for acid–base catalysis 13. Metal ions may also play an indirect role in altering the pKa of a nucleobase so it may act as a general acid or base 14,15,16,17,18,19. Protein enzymes commonly have pKa shifts of several units; it appears that RNA is capable of the same. A third strategy for catalysis is demonstrated by the glmS riboswitch. In response to high concentrations of glucosamine-6-phosphate, this riboswitch catalyses its own cleavage, which results in subsequent nuclease degradation of the mRNA. The pKas of glucosamine-6-phosphate and several analogues track closely to the observed reaction pKas of the ribozyme, suggesting a direct role in catalysis.17 Glucosamine6-phosphate binds in the active site, in position to donate an amino proton to the leaving group.20 It is also possible that the positively charged amino group stabilizes the negatively charged transition state. In either case, recruitment of a small molecule cofactor expands the functional groups available for catalysis. Aided by this cofactor, glmS can reach rates as high as 5s 2118

PROTEIN SYNTHESIS and RNA: It is necessary but not sufficient for the RNA world to be self-perpetuating: it must also have given rise to current biology featuring protein enzymes. The details of this transition are purely speculative, but it may have included the recruitment of amino acids or short peptides as enzymatic cofactors. Clearly, RNA is capable of specifically binding these ligands and could use them to assist in chemistry. Interestingly, many of the cofactors discussed above contain a nucleic acid element 20, suggesting that molecules like them may have been selected in part for their ability to bind to RNA. As peptides became more complex, the chances of encountering them randomly in solution would drop precipitously, and machinery for specifically synthesizing them would have to be developed. Eventually, this could become the mechanism for protein synthesis. Since the machinery for making proteins would have to predate proteins, the RNA world hypothesis predicts such machinery would be made of RNA. In all extant organisms, protein synthesis is carried out by a large nucleoprotein complex, the ribosome. Strikingly, the active site for peptide bond formation is composed entirely of RNA 21, 22. This is possibly the single most significant piece of data supporting the RNA world hypothesis. The ribosome manufactures proteins by linking individual amino acids through peptide bonds . As in phosphoryl transfer, this reaction may be catalysed by deprotonating the nucleophile, protonating the leaving group and stabilizing build-up of charge in the transition state. Substrate positioning is expected to play a large role, as it does for all bimolecular reactions. It is unknown (and perhaps unknowable) what a more primitive ribosome looked like, and at what point aminoacyl adenylates were used as substrates. Although a small ribozyme has been developed by in vitro selection which can form a peptide bond between two aminoacyl adenylates 23, message-directed synthesis requires a more complex system with decoding abilities. Nevertheless, the modern ribosome is a vivid demonstration that RNA has the catalytic properties necessary to begin protein synthesis. A fundamental task of proteins is to act as enzymes—catalysts that increase the rate of virtually all the chemical reactions within cells. Although RNAs are capable of catalyzing some reactions, most biological reactions are catalyzed by proteins. In the absence of enzymatic catalysis, most biochemical reactions are so slow that they would not occur under the mild conditions of temperature and pressure that are compatible with life. Enzymes accelerate the rates of such reactions by well over a million-fold, so reactions that would take years in the absence of catalysis can occur in fractions of seconds if catalyzed by the appropriate enzyme. Cells contain thousands of different enzymes, and their activities determine which of the many possible chemical reactions actually take place within the cell 24.

Properties and characteristics of protein as enzymes : Enzymes are generally globular proteins, acting alone or in larger complexes. The sequence of the amino acids specifies the structure which in turn determines the catalytic activity of the enzyme.25 Although structure determines function, a novel enzymatic activity cannot yet be predicted from structure alone.26 Enzyme structures unfold (denature) when heated or exposed to chemical denaturants and this disruption to the structure typically causes a loss of activity.27 Enzyme denaturation is normally linked to temperatures above a species' normal level; as a result, enzymes from bacteria living in volcanic environments such as hot springs are prized by industrial users for their ability to function at high temperatures, allowing enzyme-catalysed reactions to be operated at a very high rate. Enzymes are usually much larger than their substrates. Sizes range from just 62 amino acid residues, for the monomer of 4-oxalocrotonate tautomerase,28 to over 2,500 residues in the animal fatty acid synthase.29 Only a small portion of their structure (around 2–4 amino acids) is directly involved in catalysis: the catalytic site.30 This catalytic site is located next to one or more binding sites where residues orient the substrates. The catalytic site and binding site together comprise the enzyme's active site. The remaining majority of the enzyme structure serves to maintain the precise orientation and dynamics of the active site.31 Enzymes must bind their substrates before they can catalyse any chemical reaction. Enzymes are usually very specific as to what substrates they bind and then the chemical reaction catalysed. Specificity is achieved by binding pockets with complementary shape, charge and hydrophilic/hydrophobic characteristics to the substrates. Enzymes can therefore distinguish between very similar substrate molecules to be chemoselective, regioselective and stereospecific.32 Some of the enzymes showing the highest specificity and accuracy are involved in the copying and expression of the genome. Some of 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.33 This two-step process results in average error rates of less than 1 error in 100 million reactions in high-fidelity mammalian polymerases. Similar proofreading mechanisms are also found in RNA polymerase,34 aminoacyl tRNA synthetases35 and ribosomes.36 To explain the observed specificity of enzymes, in 1894 Emil Fischer proposed that both the enzyme and the substrate possess specific complementary geometric shapes that fit exactly into one another.37 This is often referred to as "the lock and key" model.1:8.3.2 This early model explains enzyme specificity, but fails to explain the stabilization of the transition state that enzymes achieve.38 In 1958, Daniel Koshland suggested a modification to the lock and key model: since enzymes are rather flexible structures, the active site is continuously reshaped by interactions with the substrate as the substrate interacts with the enzyme.39 As a result, the substrate does not simply bind to a rigid active site; the amino acid side-chains that make up the active site are molded 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.40 The active site continues to change until the substrate is completely bound, at which point the final shape and charge distribution is determined.41 Induced fit may enhance the fidelity of molecular recognition in the presence of competition and noise via the conformational proofreading mechanism.42

Enzymes can accelerate reactions in several ways, all of which lower the activation energy (ΔG‡, Gibbs free energy)43
  1. By stabilizing the transition state:
  2. Creating an environment with a charge distribution complementary to that of the transition state to lower its energy[44]
  3. By providing an alternative reaction pathway:
  4. Temporarily reacting with the substrate, forming a covalent intermediate to provide a lower energy transition state[45]
  5. By destabilising the substrate ground state:
  6. oDistorting bound substrate(s) into their transition state form to reduce the energy required to reach the transition state46
  7. By orienting the substrates into a productive arrangement to reduce the reaction entropy change47 (the contribution of this mechanism to catalysis is relatively small)48

Enzymes may use several of these mechanisms simultaneously. For example, proteases such as trypsin perform covalent catalysis using a catalytic triad, stabilise charge build-up on the transition states using an oxyanion hole, complete hydrolysis using an oriented water substrate.49,50,51,52,53

Dynamics:

Enzymes are not rigid, static structures; instead they have complex internal dynamic motions – that is, movements of parts of the enzyme's structure such as individual amino acid residues, groups of residues forming a protein loop or unit of secondary structure, or even an entire protein domain. These motions give rise to a conformational ensemble of slightly different structures that interconvert with one another at equilibrium. Different states within this ensemble may be associated with different aspects of an enzyme's function. For example, different conformations of the enzyme dihydrofolate reductase are associated with the substrate binding, catalysis, cofactor release, and product release steps of the catalytic cycle,53 consistent with catalytic resonance theory.

Types of inhibition

Competitive

A competitive inhibitor and substrate cannot bind to the enzyme at the same time.56 Often competitive inhibitors strongly resemble the real substrate of the enzyme. For example, the drug methotrexate is a competitive inhibitor of the enzyme dihydrofolate reductase, which catalyzes the reduction of dihydrofolate to tetrahydrofolate.54 The similarity between the structures of dihydrofolate and this drug are shown in the accompanying figure. This type of inhibition can be overcome with high substrate concentration. In some cases, the inhibitor can bind to a site other than the binding-site of the usual substrate and exert an allosteric effect to change the shape of the usual binding-site.57

Non-competitive

A non-competitive inhibitor binds to a site other than where the substrate binds. The substrate still binds with its usual affinity and hence Km remains the same. However the inhibitor reduces the catalytic efficiency of the enzyme so that Vmax is reduced. In contrast to competitive inhibition, non-competitive inhibition cannot be overcome with high substrate concentration.55,56–64

Uncompetitive

An uncompetitive inhibitor cannot bind to the free enzyme, only to the enzyme-substrate complex; hence, these types of inhibitors are most effective at high substrate concentration. In the presence of the inhibitor, the enzyme-substrate complex is inactive.55,64 This type of inhibition is rare.58

Mixed

A mixed inhibitor binds to an allosteric site and the binding of the substrate and the inhibitor affect each other. The enzyme's function is reduced but not eliminated when bound to the inhibitor. This type of inhibitor does not follow the Michaelis-Menten equation. 62–64

Irreversible

An irreversible inhibitor permanently inactivates the enzyme, usually by forming a covalent bond to the protein.73 Penicillin74 and aspirin75 are common drugs that act in this manner.

Functions of inhibitors

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. Major metabolic pathways such as the citric acid cycle make use of this mechanism. Since inhibitors modulate the function of enzymes they are often used as drugs. Many such drugs are reversible competitive inhibitors that resemble the enzyme's native substrate, similar to methotrexate above; other well-known examples include statins used to treat high cholesterol,62 and protease inhibitors used to treat retroviral infections such as HIV.77 A common example of an irreversible inhibitor that is used as a drug is aspirin, which inhibits the COX-1 and COX-2 enzymes that produce the inflammation messenger prostaglandin.61 Other enzyme inhibitors are poisons. For example, the poison cyanide is an irreversible enzyme inhibitor that combines with the copper and iron in the active site of the enzyme cytochrome c oxidase and blocks cellular respiration.64

 

General characteristics Catalytic RNAs: are broadly separated into two classes based on their size and reaction mechanisms (reviewed by [65-69]). The large catalytic RNAs consist of RNase P, and the group I and group II introns. These molecules range in size from a few hundred nucleotides to around 3000. They catalyze reactions that generate reaction intermediates and products with 3P hydroxyls and 5P phosphates . The small catalytic RNAs include the hammerhead, the hairpin (or paperclip), hepatitis delta and VS RNA. These molecules range in size from V35 to V155 nucleotides. They use the 2P hydroxyl of the ribose sugar as a nucleophile, and they generate products with a 2P,3P-cyclic phosphate and a 5P hydroxyl . The relationship between the size and the reaction mechanism of these molecules has raised intriguing questions about their origins and evolution. It may be that the reaction mechanism and the size of the large ribozymes are needed to bring often very distal elements of the substrate into close proximity. The small, self-cleaving, RNAs are not faced with this constraint and perhaps this permitted them to evolve smaller catalytic centers. It remains possible, however, that the relationship between the size and reaction mechanism is simply fortuitous. With one exception, all these RNAs catalyze reactions that modify themselves. Hence, they cannot be considered true enzymes or catalysts. The exception is RNase P, which processes the 5P end of tRNA precursors. It is the only known example of a naturally occurring RNA-based enzyme. However, all these molecules can be converted, with some clever engineering, into true RNA enzymes that modify other RNAs in trans without becoming altered themselves. Ribozymes increase reaction rates by up to 1011- fold and have reaction e¤ciencies, kcat/Km, up to 108 M31 min31, which is in the range for di¡usion-controlled duplex formation between oligonucleotides 65. While impressive, the rate enhancements provided by ribozymes are still V103-fold less than those provided by protein enzymes catalyzing comparable reactions 70. Moreover, ribozymes cannot compare with proteins as multiple-turnover enzymes, mostly because product release is so slow that the catalytic site of the ribozyme is easily saturated. This may be an inherent limitation of RNA enzymes, but it could also re£ect evolutionary constraints, since ribozymes generally catalyze intramolecular, single-turnover, reactions in nature. An exhaustive comparison of the enzymatic mechanistics of protein and RNA enzymes has recently been made 70. All known ribozymes have an absolute requirement for a divalent cation, which is generally Mg2‡. Some, notably within the large catalytic RNAs, require divalent cations for proper assembly of the tertiary structures as well. On this basis, catalytic RNAs are considered to be metalloenzymes, and a general two-metal-ion reaction mechanism has been proposed for the large catalytic RNAs, based on analogy with the properties of protein metalloenzymes 71.

 

CONCLUSION  

Only two reactions are observed to be catalysed by RNA in nature. However, the diversity of catalytic mechanisms for these reactions implies that RNA may be capable of much more. This is supported by the success of in vitro selection to develop ribozymes for RNA ligation 72, carbon–carbon bond formation 73, glycosidic bond formation 74and other activities. Furthermore, in vitro-selected ribozymes can use catalytic cofactors, demonstrated by an alcohol dehydrogenase 75. If the RNA world hypothesis is correct, then almost all ribozymes have been replaced by protein counterparts. Why then do some ribozymes persist in the presence of proteins, while others do not? First, the lower reaction rate could be compensated for by the increased effective concentration, since most ribozymes exist in nature as cis-acting elements, and not true trans-acting enzymes. In fact, the deficiency of many ribozymes compared with protein enzymes is not in the catalytic rate, but in binding the engineered substrate. Further, some ribozymes are mobile genetic elements, which benefit from requiring as little exogenous protein as possible (and, in fact, introns often contain their own open reading frames) 76.

 

REFERENCES 

  1. Stryer L, Berg JM, Tymoczko JL (2002). Biochemistry(5th ed.). San Francisco: W.H. Freeman. ISBN 0-7167-4955-6.
  2. Murphy JM, Farhan H, Eyers PA (2017). "Bio-Zombie: the rise of pseudoenzymes in biology". Biochem Soc Trans. 45 (2): 537–544.
  3. Murphy JM, et al. (2014). "A robust methodology to subclassify pseudokinases based on their nucleotide-binding properties". Biochemical Journal. 457 (2): 323–334.
  4. Schomburg I, Chang A, Placzek S, Söhngen C, Rother M, Lang M, Munaretto C, Ulas S, Stelzer M, Grote A, Scheer M, Schomburg D (January 2013). "BRENDA in 2013: integrated reactions, kinetic data, enzyme function data, improved disease classification: new options and contents in BRENDA". Nucleic Acids Research. 41 (Database issue): D764–72.
  5. Radzicka A, Wolfenden R (January 1995). "A proficient enzyme". Science. 267 (5194): 90–931. Bibcode:1995Sci...267...90R.
  6. Callahan BP, Miller BG (December 2007). "OMP decarboxylase—An enigma persists". Bioorganic Chemistry. 35 (6): 465–9.
  7. Gilbert, W. 1986 Origin of life: the RNA world. Nature 319, 618.
  8. Warnecke, J. M., Held, R., Busch, S. and Hartmann, R. K. 1999 Role of metal ions in the hydrolysis reaction catalyzed by RNase P RNA from Bacillus subtilis. J. Mol. Biol. 290, 433–445.
  9. Yean, S., Wuenschell, G., Termini, J. and Lin, R. 2000 Metal-ion coordination by U6 small nuclear RNA contributes to catalysis in the spliceosome. Nature 408, 881–884.
  10. Murray, J. B., Seyhan, A. A., Walter, N. G., Burke, J. M. and Scott, W. G. 1998 The hammerhead, hairpin and VS ribozymes are catalytically proficient in monovalent cations alone. Chem. Biol. 5, 587–595.
  11. Martick, M. and Scott, W. G. 2006 Tertiary contacts distant from the active site prime a ribozyme for catalysis. Cell 126, 309–320.
  12. Thomas, J. M. and Perrin, D. M. 2008 Probing general base catalysis in the hammerhead ribozyme. J. Am. Chem. Soc. 130, 15 467–15 475
  13. Das, S. R. and Piccirilli, J. A. 2005 General acid catalysis by the hepatitis delta virus ribozyme. Nat. Chem. Biol. 1, 45–52.
  14. Wilson, T. J., McLeod, A. C. and Lilley, D. M. J. 2007 A guanine nucleobase important for catalysis by the VS ribozyme. EMBO J. 26, 2489–2500.
  15. Cochrane, J. C., Lipchock, S. V. and Strobel, S. A. 2007 Structural investigation of the GlmS ribozyme bound to its catalytic cofactor. Chem. Biol. 14, 97–105.
  16. Klein, D. J. and Ferre´-D’Amare´, A. R. 2006 Structural basis of glmS ribozyme activation by glucosamine6-phosphate. Science 313, 1752–1756.
  17. Winkler, W. C., Nahvi, A., Roth, A., Collins, J. A. and Breaker, R. R. 2004 Control of gene expression by a natural metabolite-responsive ribozyme. Nature 428, 281–286
  18. Cochrane, J. C., Lipchock, S. V., Smith, K. D. and Strobel, S. A. 2009 Structural and chemical basis for glucosamine 6-phosphate binding and activation of the glmS ribozyme. Biochemistry 48, 3239–3246.
  19. The Cell: A Molecular Approach. 2nd edition. Available at : https://www.ncbi.nlm.nih.gov/books/NBK9921/ accessed [Jan 2020].
  20. White, H. B. 1976 Coenzymes as fossils of an earlier metabolic state. J. Mol. Evol. 7, 101–104.
  21. Ban, N., Nissen, P., Hansen, J., Moore, P. B. and Steitz, T. A. 2000 The complete atomic structure of the large ribosomal subunit at 2.4 A˚ resolution. Science 289, 905–920.
  22. Nissen, P., Hansen, J., Ban, N., Moore, P. B. and Steitz, T. A. 2000 The structural basis of ribosome activity in peptide bond synthesis. Science 289, 920–930.
  23. Zhang, B. and Cech, T. R. 1997 Peptide bond formation by in vitro selected ribozymes. Nature 390, 96–100.
  24. Wilson, T. J., McLeod, A. C. and Lilley, D. M. J. 2007 A guanine nucleobase important for catalysis by the VS ribozyme. EMBO J. 26, 2489–2500.
  25. Anfinsen CB (July 1973). "Principles that govern the folding of protein chains". Science. 181 (4096): 223–30. Bibcode:1973Sci...181..223A. doi:10.1126/science.181.4096.223. PMID 4124164.
  26. Dunaway-Mariano D (November 2008). "Enzyme function discovery". Structure. 16 (11): 1599–600. doi:10.1016/j.str.2008.10.001. PMID 19000810.
  27. Petsko GA, Ringe D (2003). "Chapter 1: From sequence to structure". Protein structure and function. London: New Science. p. 27. ISBN 978-1405119221.
  28. Chen LH, Kenyon GL, Curtin F, Harayama S, Bembenek ME, Hajipour G, Whitman CP (September 1992). "4-Oxalocrotonate tautomerase, an enzyme composed of 62 amino acid residues per monomer". The Journal of Biological Chemistry. 267 (25): 17716–21. PMID 1339435.
  29. Smith S (December 1994). "The animal fatty acid synthase: one gene, one polypeptide, seven enzymes". FASEB Journal. 8 (15): 1248–59. doi:10.1096/fasebj.8.15.8001737. PMID 8001737.
  30. "The Catalytic Site Atlas". The European Bioinformatics Institute. Retrieved 4 April2007.
  31. Suzuki H (2015). "Chapter 7: Active Site Structure". How Enzymes Work: From Structure to Function. Boca Raton, FL: CRC Press. pp. 117–140. ISBN 978-981-4463-92-8.
  32. Krauss G (2003). "The Regulations of Enzyme Activity". Biochemistry of Signal Transduction and Regulation (3rd ed.). Weinheim: Wiley-VCH. pp. 89–114. ISBN 9783527605767.
  33. Jaeger KE, Eggert T (August 2004). "Enantioselective biocatalysis optimized by directed evolution". Current Opinion in Biotechnology. 15 (4): 305–13. doi:10.1016/j.copbio.2004.06.007. PMID 15358000.
  34. Shevelev IV, Hübscher U (May 2002). "The 3' 5' exonucleases". Nature Reviews Molecular Cell Biology. 3 (5): 364–76. doi:10.1038/nrm804. PMID 11988770.
  35. Zenkin N, Yuzenkova Y, Severinov K (July 2006). "Transcript-assisted transcriptional proofreading". Science. 313 (5786): 518–20. Bibcode:2006Sci...313..518Z. doi:10.1126/science.1127422. PMID 16873663.
  36. Ibba M, Soll D (2000). "Aminoacyl-tRNA synthesis". Annual Review of Biochemistry. 69: 617–50. doi:10.1146/annurev.biochem.69.1.617. PMID 10966471.
  37. Rodnina MV, Wintermeyer W (2001). "Fidelity of aminoacyl-tRNA selection on the ribosome: kinetic and structural mechanisms". Annual Review of Biochemistry. 70: 415–35. doi:10.1146/annurev.biochem.70.1.415. PMID 11395413.
  38. Khersonsky O, Tawfik DS (2010). "Enzyme promiscuity: a mechanistic and evolutionary perspective". Annual Review of Biochemistry. 79: 471–505. doi:10.1146/annurev-biochem-030409-143718. PMID 20235827.
  39. O'Brien PJ, Herschlag D (April 1999). "Catalytic promiscuity and the evolution of new enzymatic activities". Chemistry and Biology. 6 (4): R91–R105. doi:10.1016/S1074-5521(99)80033-7. PMID 10099128.
  40.  Fischer E (1894). "Einfluss der Configuration auf die Wirkung der Enzyme" [Influence of configuration on the action of enzymes]. Berichte der Deutschen Chemischen Gesellschaft zu Berlin (in German). 27 (3): 2985–93. doi:10.1002/cber.18940270364. From page 2992: "Um ein Bild zu gebrauchen, will ich sagen, dass Enzym und Glucosid wie Schloss und Schlüssel zu einander passen müssen, um eine chemische Wirkung auf einander ausüben zu können." (To use an image, I will say that an enzyme and a glucoside [i.e., glucose derivative] must fit like a lock and key, in order to be able to exert a chemical effect on each other.)
  41. Cooper GM (2000). "Chapter 2.2: The Central Role of Enzymes as Biological Catalysts". The Cell: a Molecular Approach (2nd ed.). Washington (DC ): ASM Press. ISBN 0-87893-106-6.
  42. Koshland DE (February 1958). "Application of a Theory of Enzyme Specificity to Protein Synthesis". Proceedings of the National Academy of Sciences of the United States of America. 44 (2): 98–104. Bibcode:1958PNAS...44...98K. doi:10.1073/pnas.44.2.98. PMC 335371. PMID 16590179.
  43. Vasella A, Davies GJ, Böhm M (October 2002). "Glycosidase mechanisms". Current Opinion in Chemical Biology. 6 (5): 619–29. doi:10.1016/S1367-5931(02)00380-0. PMID 12413546.
  44. Boyer R (2002). "Chapter 6: Enzymes I, Reactions, Kinetics, and Inhibition". Concepts in Biochemistry (2nd ed.). New York, Chichester, Weinheim, Brisbane, Singapore, Toronto.: John Wiley and Sons, Inc. pp. 137–8. ISBN 0-470-00379-0. OCLC 51720783.
  45. Savir Y, Tlusty T (2007). Scalas E (ed.). "Conformational proofreading: the impact of conformational changes on the specificity of molecular recognition" (PDF). PLoS ONE. 2(5): e468. Bibcode:2007PLoSO...2..468S. doi:10.1371/journal.pone.0000468. PMC 1868595. PMID 17520027. Archived from the original (PDF) on 14 May 2011. Retrieved 22 August 2010.
  46. Fersht A (1985). Enzyme Structure and Mechanism. San Francisco: W.H. Freeman. pp. 50–2. ISBN 978-0-7167-1615-0.
  47. Warshel A, Sharma PK, Kato M, Xiang Y, Liu H, Olsson MH (August 2006). "Electrostatic basis for enzyme catalysis". Chemical Reviews. 106 (8): 3210–35. doi:10.1021/cr0503106. PMID 16895325.
  48. Cox MM, Nelson DL (2013). "Chapter 6.2: How enzymes work". Lehninger Principles of Biochemistry (6th ed.). New York, N.Y.: W.H. Freeman. p. 195. ISBN 978-1464109621.
  49. Benkovic SJ, Hammes-Schiffer S (August 2003). "A perspective on enzyme catalysis". Science. 301 (5637): 1196–202. Bibcode:2003Sci...301.1196B. doi:10.1126/science.1085515. PMID 12947189.
  50. Jencks WP (1987). Catalysis in Chemistry and Enzymology. Mineola, N.Y: Dover. ISBN 978-0-486-65460-7.
  51. Villa J, Strajbl M, Glennon TM, Sham YY, Chu ZT, Warshel A (October 2000). "How important are entropic contributions to enzyme catalysis?". Proceedings of the National Academy of Sciences of the United States of America. 97 (22): 11899–904. Bibcode:2000PNAS...9711899V. doi:10.1073/pnas.97.22.11899. PMC 17266. PMID 11050223.
  52.  Polgár, L. (7 July 2005). "The catalytic triad of serine peptidases". Cellular and Molecular Life Sciences. 62 (19–20): 2161–2172. doi:10.1007/s00018-005-5160-x. ISSN 1420-682X. PMID 16003488.
  53. Ramanathan A, Savol A, Burger V, Chennubhotla CS, Agarwal PK (2014). "Protein conformational populations and functionally relevant substates". Acc. Chem. Res. 47 (1): 149–56. doi:10.1021/ar400084s. PMID 23988159.
  54. Goodsell, David S. (1 August 1999). "The Molecular Perspective: Methotrexate". The Oncologist. 4 (4): 340–341. ISSN 1083-7159. PMID 10476546.
  55.  Cornish-Bowden A (2004). Fundamentals of Enzyme Kinetics (3 ed.). London: Portland Press. ISBN 1-85578-158-1.
  56. Price NC (1979). "What is meant by 'competitive inhibition'?". Trends in Biochemical Sciences. 4 (11): N272–N273. doi:10.1016/0968-0004(79)90205-6.
  57. Wu, Peng; Clausen, Mads H.; Nielsen, Thomas E. (1 December 2015). "Allosteric small-molecule kinase inhibitors" (PDF). Pharmacology and Therapeutics. 156: 59–68. doi:10.1016/j.pharmthera.2015.10.002. ISSN 0163-7258. PMID 26478442.
  58. Cornish-Bowden A (July 1986). "Why is uncompetitive inhibition so rare? A possible explanation, with implications for the design of drugs and pesticides". FEBS Letters. 203(1): 3–6. doi:10.1016/0014-5793(86)81424-7. PMID 3720956.
  59. Strelow, John M. (1 January 2017). "A Perspective on the Kinetics of Covalent and Irreversible Inhibition". SLAS DISCOVERY: Advancing Life Sciences RandD. 22 (1): 3–20. doi:10.1177/1087057116671509. ISSN 2472-5552. PMID 27703080.
  60. Fisher JF, Meroueh SO, Mobashery S (February 2005). "Bacterial resistance to beta-lactam antibiotics: compelling opportunism, compelling opportunity". Chemical Reviews. 105 (2): 395–424. doi:10.1021/cr030102i. PMID 15700950.
  61. Johnson DS, Weerapana E, Cravatt BF (June 2010). "Strategies for discovering and derisking covalent, irreversible enzyme inhibitors". Future Medicinal Chemistry. 2 (6): 949–64. doi:10.4155/fmc.10.21. PMC 2904065. PMID 20640225.
  62. Endo A (1 November 1992). "The discovery and development of HMG-CoA reductase inhibitors" (PDF). J. Lipid Res. 33 (11): 1569–82. PMID 1464741.
  63. Wlodawer A, Vondrasek J (1998). "Inhibitors of HIV-1 protease: a major success of structure-assisted drug design". Annual Review of Biophysics and Biomolecular Structure. 27: 249–84. doi:10.1146/annurev.biophys.27.1.249. PMID 9646869.
  64. Yoshikawa S, Caughey WS (May 1990). "Infrared evidence of cyanide binding to iron and copper sites in bovine heart cytochrome c oxidase. Implications regarding oxygen reduction". The Journal of Biological Chemistry. 265 (14): 7945–58. PMID 2159465.
  65. Cech, T.R. (1993) Structure and mechanism of the large catalytic RNAs: group I and group II introns and ribonuclease P. In: The RNA World (Gesteland, R.F. and Atkins, J.F., Eds.), pp. 239^269. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.
  66. Symons, R.H. (1994) Ribozymes. Curr. Opin. Struct. Biol. 4, 322^330.
  67. Scott, W.G. and Klug, A. (1996) Ribozymes: structure and mechanism in RNA catalysis. Trends Biochem. Sci. 21, 220^ 224.
  68. Symons, R.H. (1997) Plant pathogenic RNAs and RNA catalysis. Nucleic Acids Res. 25, 2683^2689.
  69. Tanner, N.K. (1998) Ribozymes: caracte¨ristiques et applications. Virologie 2, 127^137.
  70. Narlikar, G.J. and Herschlag, D. (1997) Mechanistic aspects of enzymatic catalysis: lessons from comparison of RNA and protein enzymes. Annu. Rev. Biochem. 66, 19^59.
  71. Steitz, T.A. and Steitz, J.A. (1993) A general two-metal-ion mechanism for catalytic RNA. Proc. Natl. Acad. Sci. USA 90, 6498^6502.
  72. Bartel, D. and Szostak, J. 1993 Isolation of new ribozymes from a large pool of random sequences (see comment). Science 261, 1411–1418.
  73. Tarasow, T. M., Tarasow, S. L. and Eaton, B. E. 1997 RNA-catalysed carbon–carbon bond formation. Nature 389, 54–57. (doi:10.1038/37950)
  74. Unrau, P. J. and Bartel, D. P. 1998 RNA-catalysed nucleotide synthesis. Nature 395, 260–263.
  75. Tsukiji, S., Pattnaik, S. B. and Suga, H. 2003 An alcohol dehydrogenase ribozyme. Nat. Struct. Biol. 10, 713–717.
  76. Lambowitz, A. M. and Belfort, M. 1993 Introns as mobile genetic elements. Annu. Rev. Biochem. 62, 587–622.

Policy for Articles with Open Access
Authors who publish with MedPulse International Journal of Biochemistry (Print ISSN:2550-763X) (Online ISSN: 2636-4573) agree to the following terms:
Authors retain copyright and grant the journal right of first publication with the work simultaneously licensed under a Creative Commons Attribution License that allows others to share the work with an acknowledgement of the work's authorship and initial publication in this journal.
Authors are permitted and encouraged to post links to their work online (e.g., in institutional repositories or on their website) prior to and during the submission process, as it can lead to productive exchanges, as well as earlier and greater citation of published work.