Morpheein

Last updated
Proteins that function as morpheeins are illustrated using a dice analogy where one dice can morph into two different shapes, cubic and tetrahedral. The illustrated assemblies apply a rule that the dice face with one spot must contact the dice face with four spots. To satisfy the rule for each dice in an assembly, the cubic dice can only form a tetramer and the tetrahedral dice can only assemble to a pentamer. This is analogous to two different conformations (morpheein forms) of a protein subunit each dictating assembly to a different oligomer. All dice in one assembly must be of the same shape before assembly. Thus, for example, the tetramer must come apart, and its component dice must change shape to a pyramid before they can participate in assembly into a pentamer. Morpheein dice.PNG
Proteins that function as morpheeins are illustrated using a dice analogy where one dice can morph into two different shapes, cubic and tetrahedral. The illustrated assemblies apply a rule that the dice face with one spot must contact the dice face with four spots. To satisfy the rule for each dice in an assembly, the cubic dice can only form a tetramer and the tetrahedral dice can only assemble to a pentamer. This is analogous to two different conformations (morpheein forms) of a protein subunit each dictating assembly to a different oligomer. All dice in one assembly must be of the same shape before assembly. Thus, for example, the tetramer must come apart, and its component dice must change shape to a pyramid before they can participate in assembly into a pentamer.

Morpheeins are proteins that can form two or more different homo-oligomers (morpheein forms), but must come apart and change shape to convert between forms. The alternate shape may reassemble to a different oligomer. The shape of the subunit dictates which oligomer is formed. [1] [2] Each oligomer has a finite number of subunits (stoichiometry). Morpheeins can interconvert between forms under physiological conditions and can exist as an equilibrium of different oligomers. These oligomers are physiologically relevant and are not misfolded protein; this distinguishes morpheeins from prions and amyloid. The different oligomers have distinct functionality. Interconversion of morpheein forms can be a structural basis for allosteric regulation, an idea noted many years ago, [3] [4] and later revived. [1] [2] [5] [6] A mutation that shifts the normal equilibrium of morpheein forms can serve as the basis for a conformational disease. [7] Features of morpheeins can be exploited for drug discovery. [1] [5] [8] The dice image (Fig 1) represents a morpheein equilibrium containing two different monomeric shapes that dictate assembly to a tetramer or a pentamer. The one protein that is established to function as a morpheein is porphobilinogen synthase, [2] [9] [10] though there are suggestions throughout the literature that other proteins may function as morpheeins (for more information see "Table of Putative Morpheeins" below).

Contents

Implications for drug discovery

Conformational differences between subunits of different oligomers and related functional differences of a morpheein provide a starting point for drug discovery. Protein function is dependent on the oligomeric form; therefore, the protein's function can be regulated by shifting the equilibrium of forms. A small molecule compound can shift the equilibrium either by blocking or favoring formation of one of the oligomers. The equilibrium can be shifted using a small molecule that has a preferential binding affinity for only one of the alternate morpheein forms. An inhibitor of porphobilinogen synthase with this mechanism of action has been documented. [5]


Implications for allosteric regulation

The morpheein model of allosteric regulation has similarities to and differences from other models. [1] [6] [11] The concerted model (the Monod, Wyman and Changeux (MWC) model) of allosteric regulation requires all subunits to be in the same conformation or state within an oligomer like the morpheein model. [12] [13] However, neither this model nor the sequential model (Koshland, Nemethy, and Filmer model) takes into account that the protein may dissociate to interconvert between oligomers. [12] [13] [14] [15] Nonetheless, shortly after these theories were described, two groups of workers [3] [4] proposed what is now called the morpheein model and showed that it accounted for the regulatory behavior of glutamate dehydrogenase. [16] Kurganov and Friedrich discussed models of this kind extensively in their books. [17] [18]

Implications for teaching about protein structure-function relationships

It is generally taught [ citation needed ] that a given amino acid sequence will have only one physiologically relevant (native) quaternary structure; morpheeins challenge this concept. The morpheein model does not require gross changes in the basic protein fold. [1] The conformational differences that accompany conversion between oligomers may be similar to the protein motions necessary for function of some proteins. [19] The morpheein model highlights the importance of conformational flexibility for protein functionality and offers a potential explanation for proteins showing non-Michaelis-Menten kinetics, hysteresis, and/or protein concentration dependent specific activity. [11]

Implications for understanding the structural basis for disease

The term "conformational disease" generally encompasses mutations that result in misfolded proteins that aggregate, such as Alzheimer's and Creutzfeldt–Jakob diseases. [20] In light of the discovery of morpheeins, however, this definition could be expanded to include mutations that shift an equilibrium of alternate oligomeric forms of a protein. An example of such a conformational disease is ALAD porphyria, which results from a mutation of porphobilinogen synthase that causes a shift in its morpheein equilibrium. [7]

Table of proteins whose published behavior is consistent with that of a morpheein

[6]

ProteinExample speciesEC numberCAS numberAlternate oligomersEvidence
Acetyl-CoA carboxylase-1Gallus domesticus EC 6.4.1.2 9023-93-2 inactive dimer, active dimer, larger [21] Effector molecules impact multimerization, [22] Multiple/protein moonlighting functions [21]
α-AcetylgalactosaminidaseBos taurus EC 4.3.2.2 9027-81-0 inactive monomer, active tetramer [23] Substrate binding/turnover impacts multimerization, [23] Protein concentration dependent specific activity, [24] Different assemblies have different activities, [24] Conformationally distinct oligomeric forms. [23] [24]
Adenylosuccinate lyase Bacillus subtilis EC 4.3.2.2 9027-81-0 monomer, dimer, trimer, tetramer [25] Mutations shift the equilibrium of oligomers, [26] Oligomer-dependent kinetic parameters, [26] Protein concentration dependent molecular weight [26]
Aristolochene synthase Penicillium roqueforti EC 4.2.3.9 94185-89-4 monomer, higher order [27] Protein concentration dependent specific activity [28]
L-Asparaginase Leptosphaeria michotii EC 3.5.1.1 9015-68-3 dimer, tetramer, inactive octamer [29] Substrate binding/turnover impacts multimerization [30]
Aspartokinase Escherichia coli EC 2.7.2.4 & EC 1.1.1.3 9012-50-4 monomer, dimer, tetramer [31] [32] Multiple/protein moonlighting functions, [33] Conformationally distinct oligomeric forms [32]
ATPase of the ABCA1 transporterHomo sapiensdimer, tetramer [34] Substrate binding/turnover impacts multimerization [34]
Biotin—(acetyl-CoA-carboxylase) ligase holoenzyme synthetaseEscherichia coli EC 6.3.4.15 37340-95-7 monomer, dimer [35] Multiple/protein moonlighting functions, [35] Different assemblies have different activities [36]
Chorismate mutase Escherichia coli EC 5.4.99.5 9068-30-8 dimer, trimer, hexamerConformationally distinct oligomeric forms [37]
Citrate synthase Escherichia coli EC 2.3.3.1 9027-96-7 monomer, dimer, trimer, tetramer, pentamer, hexamer, dodecamer [38] Substrate binding/turnover impacts multimerization, [38] Characterized equilibrium of oligomers, [38] Protein concentration dependent specific activity, [38] pH-dependent oligomeric equilibrium [38]
Cyanovirin-N Nostoc ellipsosporum 918555-82-5 monomer and domain-swapped dimer [39] [40] Characterized equilibrium of oligomers, [41] [42] Conformationally distinct oligomeric forms [41] [42]
3-oxoacid CoA-transferase Sus scrofa domestica EC 2.8.3.5 9027-43-4 dimer, tetramer [43] Chromatographically separable oligomers, [43] Substrate might preferentially stabilize one form [43]
Cystathionine β-synthase Homo sapiens EC 4.2.1.22 9023-99-8 multiple - ranges from dimer to 16-mer [44] Effector molecules impact multimerization, [45] Mutations shift the equilibrium of oligomers, [46] Different assemblies have different activities, [45] disease-causing mutations at sites distant from active site [47]
D-amino acid oxidase EC 1.4.3.3 9000-88-8 monomers, dimers, higher-order oligomers [48] [49] Oligomer-dependent kinetic parameters [48] [49]
Dihydrolipoamide dehydrogenase Sus scrofa domestica EC 1.8.1.4 9001-18-7 monomer, two different dimer forms, tetramer [50] Multiple/protein moonlighting functions, [50] Different assemblies have different activities, [50] pH-dependent oligomeric equilibrium, [50] Conformationally distinct oligomeric forms [51] [52] [53]
Dopamine β-monooxygenase Bos taurus EC 1.14.17.1 9013-38-1 dimers, tetramers [54] [55] [56] Effector molecules impact multimerization, [54] [55] [56] Characterized equilibrium of oligomers, [54] [55] [56] Oligomer-dependent kinetic parameters [54] [55] [56]
Geranylgeranyl pyrophosphate synthase / Farnesyltranstransferase Homo sapiens EC 2.5.1.29 9032-58-0 hexamer, octamer [57] [58] [59] Effector molecules impact multimerization [58]
GDP-mannose 6-dehydrogenase Pseudomonas aeruginosa EC 1.1.1.132 37250-63-8 trimer, 2 tetramers, and hexamer [60] [61] Protein concentration dependent specific activity, [62] Kinetic hysteresis [62]
Glutamate dehydrogenase Bos taurus EC 1.4.1.2 9001-46-1 active & inactive hexamers, higher order [63] Effector molecules impact multimerization, [64] Characterized equilibrium of oligomers [63]
Glutamate racemase Mycobacterium tuberculosis, Escherichia coli, Bacillus subtilis, Aquifex pyrophilus EC 5.1.1.3 9024-08-02 monomer, 2 dimers, tetramer [65] [66] [67] [68] [69] Multiple/protein moonlighting functions, [70] [71] [72] Characterized equilibrium of oligomers, [68] [69] Conformationally distinct oligomeric forms [65] [66] [67]
Glyceraldehyde-3-phosphate dehydrogenase Oryctolagus cuniculas, Sus scrofa domestica EC 1.2.1.12 1.2.1.12 9001-50-7 monomer, dimer, tetramer [73] Characterized equilibrium of oligomers, [74] Different assemblies have different activities [75]
Glycerol kinase Escherichia coli EC 2.7.1.30 9030-66-4 monomer and 2 tetramers [76] [77] [78] Characterized equilibrium of oligomers, [76] [77] [78] [79] Conformationally distinct oligomeric forms, [79] [80] Effector functions by preventing domain motion [80]
HIV-Integrase Human immunodeficiency virus-1 EC 2.7.7.- monomer, dimer, tetramer, higher order [81] [82] [83] Effector molecules impact multimerization, [84] Multiple/protein moonlighting functions, [81] [82] [83] Different assemblies have different activities [83] [84]
HPr-Kinase/phosphataseBacillus subtilis, Lactobacillus casei, Mycoplasma pneumoniae, Staphylococcus xylosus EC 2.7.1.-/ EC 3.1.3.- 9026-43-1 monomers, dimers, trimers, hexamers [85] [86] [87] [88] [89] [90] Effector molecules impact multimerization, [89] Multiple/protein moonlighting functions, [89] Different assemblies have different activities, [89] pH-dependent oligomeric equilibrium [89]
Lactate dehydrogenase Bacillus stearothermophilus EC 1.1.1.27 9001-60-9 2 dimers, tetramer [91] [92] Effector molecules impact multimerization, [92] Characterized equilibrium of oligomers, [92] Protein concentration dependent specific activity, [92] Mutations shift the equilibrium of oligomers, [93] Oligomer-dependent kinetic parameters, [92] Conformationally distinct oligomeric forms [94]
Lon proteaseEscherichia coli, Mycobacterium smegmatis EC 3.4.21.53 79818-35-2 monomer, dimer, trimer, tetramer [95] [96] Effector molecules impact multimerization, [95] [96] Substrate binding/turnover impacts multimerization, [95] [96] Protein concentration dependent specific activity, [97] Kinetic hysteresis [97]
Mitochondrial NAD(P)+ Malic enzyme / [[malate dehydrogenase (oxaloacetate-decarboxylating) (NADP+)]]Homo sapiens EC 1.1.1.40 9028-47-1 monomer, 2 dimers, tetramer [98] [99] Effector molecules impact multimerization, [98] Mutations shift the equilibrium of oligomers, [100] Kinetic hysteresis, [99]
Peroxiredoxins Salmonella typhimurium EC 1.6.4.- & EC 1.11.1.15 207137-51-7 2 dimers, decamerConformationally distinct oligomeric forms, [101] Different assemblies have different activities [102]
Phenylalanine hydroxylase Homo sapiens EC 1.14.16.1 9029-73-6 high activity tetramer, low activity tetramer [103] Substrate binding/turnover impacts multimerization, [104] [105] Conformationally distinct oligomeric forms [106] [107]
Phosphoenolpyruvate carboxylase Escherichia coli, Zea mays EC 9067-77-0 inactive dimer, active tetramer [108] Effector molecules impact multimerization, Characterized equilibrium of oligomers, [108] Kinetic hysteresis, [108] Conformationally distinct oligomeric forms [109]
Phosphofructokinase Bacillus stearothermophilus, Thermus thermophilus EC 2.7.1.11 9001-80-3 inactive dimer, active tetramer [108] [110] Effector molecules impact multimerization, [108] [110] Characterized equilibrium of oligomers [108] [110]
Polyphenol oxidase Agaricus bisporus, Malus domestica, Lactuca sativa L. EC 1.10.3.1 9002-10-2 monomer, trimer, tetramer, octamer, dodecamer [111] [112] Multiple/protein moonlighting functions, [113] Substrate binding/turnover impacts multimerization, [114] Different assemblies have different activities, [115] Kinetic hysteresis [114]
Porphobilinogen synthase Drosophila melanogaster, Danio rerio EC 4.2.1.24 9036-37-7 dimer, hexamer, octamer [116] [117] PBGS is the prototype morpheein. [116]
Pyruvate kinase Homo sapiens EC 2.7.1.40 9001-59-6 active and inactive dimers, active tetramer, monomer, trimer, pentamer [118] [119] Conformationally distinct oligomeric forms [118] [119]
Ribonuclease A Bos taurus EC 3.1.27.5 3.1.27.5 9901-99-4 monomer, dimer, trimer, tetramer, hexamer, pentamer, higher order [120] [121] [122] [123] [124] Multiple/protein moonlighting functions, [125] [126] [127] Different assemblies have different activities, [125] [126] [127] Conformationally distinct oligomeric forms [121] [123] [124]
Ribonucleotide reductase Mus musculus EC 1.17.4.1 9047-64-7 tetramer, hexamer [128] [129] [130] [131] Effector molecules impact multimerization [131]
S-adenosyl-L-homocysteine hydrolase Dictyostelium discoideum EC 3.3.1.1 9025-54-1 tetramer and other [132] [133] [134] Effector molecules impact multimerization [132]
Biodegrative threonine dehydratase / threonine ammonia-lyase Escherichia coli EC 4.3.1.19 4.3.1.19 774231-81-1 2 monomers, 2 tetramers [135] [136] [137] Effector molecules impact multimerization, [137] Characterized equilibrium of oligomers, [135] [136] Different assemblies have different activities [135] [136] [137]
β-Tryptase Homo sapiens EC 3.4.21.59 97501-93-4 active and inactive monomers, active and inactive tetramers [138] [139] [140] [141] [142] [143] [144] [145] [146] [147] Protein concentration dependent specific activity, [148] Characterized equilibrium of oligomers [148]
Tumor necrosis factor-α Homo sapiens 94948-61-5 monomer, dimer, trimer [149] [150] Different assemblies have different activities [151]
Uracil phosphoribosyltransferase Escherichia coli EC 2.4.2.9 9030-24-4 trimer, pentamer [152] Effector molecules impact multimerization, [152] Substrate binding/turnover impacts multimerization, [152] Different assemblies have different activities [152]

Related Research Articles

<span class="mw-page-title-main">Coenzyme A</span> Coenzyme, notable for its synthesis and oxidation role

Coenzyme A (CoA, SHCoA, CoASH) is a coenzyme, notable for its role in the synthesis and oxidation of fatty acids, and the oxidation of pyruvate in the citric acid cycle. All genomes sequenced to date encode enzymes that use coenzyme A as a substrate, and around 4% of cellular enzymes use it (or a thioester) as a substrate. In humans, CoA biosynthesis requires cysteine, pantothenate (vitamin B5), and adenosine triphosphate (ATP).

<span class="mw-page-title-main">Allosteric regulation</span> Regulation of enzyme activity

In biochemistry, allosteric regulation is the regulation of an enzyme by binding an effector molecule at a site other than the enzyme's active site.

<span class="mw-page-title-main">Uridine monophosphate synthase</span> Protein-coding gene in the species Homo sapiens

The enzyme Uridine monophosphate synthase catalyses the formation of uridine monophosphate (UMP), an energy-carrying molecule in many important biosynthetic pathways. In humans, the gene that codes for this enzyme is located on the long arm of chromosome 3 (3q13).

<span class="mw-page-title-main">Tryptophan synthase</span>

Tryptophan synthase or tryptophan synthetase is an enzyme that catalyses the final two steps in the biosynthesis of tryptophan. It is commonly found in Eubacteria, Archaebacteria, Protista, Fungi, and Plantae. However, it is absent from Animalia. It is typically found as an α2β2 tetramer. The α subunits catalyze the reversible formation of indole and glyceraldehyde-3-phosphate (G3P) from indole-3-glycerol phosphate (IGP). The β subunits catalyze the irreversible condensation of indole and serine to form tryptophan in a pyridoxal phosphate (PLP) dependent reaction. Each α active site is connected to a β active site by a 25 angstrom long hydrophobic channel contained within the enzyme. This facilitates the diffusion of indole formed at α active sites directly to β active sites in a process known as substrate channeling. The active sites of tryptophan synthase are allosterically coupled.

<span class="mw-page-title-main">Glutamate dehydrogenase</span> Hexameric enzyme

Glutamate dehydrogenase is an enzyme observed in both prokaryotes and eukaryotic mitochondria. The aforementioned reaction also yields ammonia, which in eukaryotes is canonically processed as a substrate in the urea cycle. Typically, the α-ketoglutarate to glutamate reaction does not occur in mammals, as glutamate dehydrogenase equilibrium favours the production of ammonia and α-ketoglutarate. Glutamate dehydrogenase also has a very low affinity for ammonia, and therefore toxic levels of ammonia would have to be present in the body for the reverse reaction to proceed. However, in brain, the NAD+/NADH ratio in brain mitochondria encourages oxidative deamination. In bacteria, the ammonia is assimilated to amino acids via glutamate and aminotransferases. In plants, the enzyme can work in either direction depending on environment and stress. Transgenic plants expressing microbial GLDHs are improved in tolerance to herbicide, water deficit, and pathogen infections. They are more nutritionally valuable.

<span class="mw-page-title-main">Malate dehydrogenase</span> Class of enzymes

Malate dehydrogenase (EC 1.1.1.37) (MDH) is an enzyme that reversibly catalyzes the oxidation of malate to oxaloacetate using the reduction of NAD+ to NADH. This reaction is part of many metabolic pathways, including the citric acid cycle. Other malate dehydrogenases, which have other EC numbers and catalyze other reactions oxidizing malate, have qualified names like malate dehydrogenase (NADP+).

<span class="mw-page-title-main">Tryptase</span> Class of enzymes

Tryptase is the most abundant secretory granule-derived serine proteinase contained in mast cells and has been used as a marker for mast cell activation. Club cells contain tryptase, which is believed to be responsible for cleaving the hemagglutinin surface protein of influenza A virus, thereby activating it and causing the symptoms of flu.

<span class="mw-page-title-main">Purine nucleoside phosphorylase</span> Enzyme

Purine nucleoside phosphorylase, PNP, PNPase or inosine phosphorylase is an enzyme that in humans is encoded by the NP gene. It catalyzes the chemical reaction

<span class="mw-page-title-main">Phosphoribosyl pyrophosphate</span> Chemical compound

Phosphoribosyl pyrophosphate (PRPP) is a pentose phosphate. It is a biochemical intermediate in the formation of purine nucleotides via inosine-5-monophosphate, as well as in pyrimidine nucleotide formation. Hence it is a building block for DNA and RNA. The vitamins thiamine and cobalamin, and the amino acid tryptophan also contain fragments derived from PRPP. It is formed from ribose 5-phosphate (R5P) by the enzyme ribose-phosphate diphosphokinase:

<span class="mw-page-title-main">CTP synthetase</span> Enzyme

CTP synthase is an enzyme involved in pyrimidine biosynthesis that interconverts UTP and CTP.

<span class="mw-page-title-main">Dihydrolipoamide dehydrogenase</span> Protein-coding gene in the species Homo sapiens

Dihydrolipoamide dehydrogenase (DLD), also known as dihydrolipoyl dehydrogenase, mitochondrial, is an enzyme that in humans is encoded by the DLD gene. DLD is a flavoprotein enzyme that oxidizes dihydrolipoamide to lipoamide.

In enzymology, a GDP-mannose 6-dehydrogenase (EC 1.1.1.132) is an enzyme that catalyzes the chemical reaction

Malate dehydrogenase (oxaloacetate-decarboxylating) (NADP<sup>+</sup>) Enzyme

Malate dehydrogenase (oxaloacetate-decarboxylating) (NADP+) (EC 1.1.1.40) or NADP-malic enzyme (NADP-ME) is an enzyme that catalyzes the chemical reaction in the presence of a bivalent metal ion:

<span class="mw-page-title-main">Delta-aminolevulinic acid dehydratase</span> Protein-coding gene in the species Homo sapiens

Aminolevulinic acid dehydratase (porphobilinogen synthase, or ALA dehydratase, or aminolevulinate dehydratase) is an enzyme (EC 4.2.1.24) that in humans is encoded by the ALAD gene. Porphobilinogen synthase (or ALA dehydratase, or aminolevulinate dehydratase) synthesizes porphobilinogen through the asymmetric condensation of two molecules of aminolevulinic acid. All natural tetrapyrroles, including hemes, chlorophylls and vitamin B12, share porphobilinogen as a common precursor. Porphobilinogen synthase is the prototype morpheein.

<span class="mw-page-title-main">Chorismate mutase</span>

In enzymology, chorismate mutase is an enzyme that catalyzes the chemical reaction for the conversion of chorismate to prephenate in the pathway to the production of phenylalanine and tyrosine, also known as the shikimate pathway. Hence, this enzyme has one substrate, chorismate, and one product, prephenate. Chorismate mutase is found at a branch point in the pathway. The enzyme channels the substrate, chorismate to the biosynthesis of tyrosine and phenylalanine and away from tryptophan. Its role in maintaining the balance of these aromatic amino acids in the cell is vital. This is the single known example of a naturally occurring enzyme catalyzing a pericyclic reaction. Chorismate mutase is only found in fungi, bacteria, and higher plants. Some varieties of this protein may use the morpheein model of allosteric regulation.

<span class="mw-page-title-main">3-oxoacid CoA-transferase</span> Enzyme family

In enzymology, a 3-oxoacid CoA-transferase is an enzyme that catalyzes the chemical reaction

<span class="mw-page-title-main">PDK2</span> Protein-coding gene in the species Homo sapiens

Pyruvate dehydrogenase kinase isoform 2 (PDK2) also known as pyruvate dehydrogenase lipoamide kinase isozyme 2, mitochondrial is an enzyme that in humans is encoded by the PDK2 gene. PDK2 is an isozyme of pyruvate dehydrogenase kinase.

<span class="mw-page-title-main">Malate dehydrogenase 2</span> Enzyme that oxidizes malate to oxaloacetate in Krebs cycle

Malate dehydrogenase, mitochondrial also known as malate dehydrogenase 2 is an enzyme that in humans is encoded by the MDH2 gene.

<span class="mw-page-title-main">Dihydrodipicolinate synthase</span> Class of enzymes

4-Hydroxy-tetrahydrodipicolinate synthase (EC 4.3.3.7, dihydrodipicolinate synthase, dihydropicolinate synthetase, dihydrodipicolinic acid synthase, L-aspartate-4-semialdehyde hydro-lyase (adding pyruvate and cyclizing), dapA (gene)) is an enzyme with the systematic name L-aspartate-4-semialdehyde hydro-lyase (adding pyruvate and cyclizing; (4S)-4-hydroxy-2,3,4,5-tetrahydro-(2S)-dipicolinate-forming). This enzyme catalyses the following chemical reaction

Edith Wilson Miles is a biochemist known for her work on the structure and function of enzymes, especially her work on tryptophan synthase.

References

  1. 1 2 3 4 5 Jaffe, Eileen K. (2005). "Morpheeins – a new structural paradigm for allosteric regulation". Trends in Biochemical Sciences. 30 (9): 490–7. doi:10.1016/j.tibs.2005.07.003. PMID   16023348.
  2. 1 2 3 Breinig, Sabine; Kervinen, Jukka; Stith, Linda; Wasson, Andrew S; Fairman, Robert; Wlodawer, Alexander; Zdanov, Alexander; Jaffe, Eileen K (2003). "Control of tetrapyrrole biosynthesis by alternate quaternary forms of porphobilinogen synthase". Nature Structural Biology. 10 (9): 757–63. doi:10.1038/nsb963. PMID   12897770. S2CID   24188785.
  3. 1 2 Frieden, C (1967). "Treatment of enzyme kinetic data. II. The multisite case: comparison of allosteric models and a possible new mechanism". J. Biol. Chem. 242 (18): 4045–4052. doi: 10.1016/S0021-9258(18)95776-5 . PMID   6061697.
  4. 1 2 Nichol, L W; Jackson, W J H; Winzor, D J (1967). "A theoretical study of the binding of small molecules to a polymerizing protein system: a model for allosteric effects". Biochemistry. 6 (8): 2449–2456. doi:10.1021/bi00860a022. PMID   6049469.
  5. 1 2 3 Lawrence, Sarah H.; Ramirez, Ursula D.; Tang, Lei; Fazliyez, Farit; Kundrat, Lenka; Markham, George D.; Jaffe, Eileen K. (2008). "Shape Shifting Leads to Small-Molecule Allosteric Drug Discovery". Chemistry & Biology. 15 (6): 586–96. doi:10.1016/j.chembiol.2008.04.012. PMC   2703447 . PMID   18559269.
  6. 1 2 3 Selwood, Trevor; Jaffe, Eileen K. (2012). "Dynamic dissociating homo-oligomers and the control of protein function". Archives of Biochemistry and Biophysics. 519 (2): 131–43. doi:10.1016/j.abb.2011.11.020. PMC   3298769 . PMID   22182754.
  7. 1 2 Jaffe, Eileen K.; Stith, Linda (2007). "ALAD Porphyria is a Conformational Disease". The American Journal of Human Genetics. 80 (2): 329–37. doi:10.1086/511444. PMC   1785348 . PMID   17236137.
  8. Jaffe, Eileen K. (2010). "Morpheeins – A New Pathway For Allosteric Drug Discovery". The Open Conference Proceedings Journal. 1: 1–6. doi: 10.2174/2210289201001010001 (inactive 2024-03-28). PMC   3107518 . PMID   21643557.{{cite journal}}: CS1 maint: DOI inactive as of March 2024 (link)
  9. Tang, L.; Stith, L; Jaffe, EK (2005). "Substrate-induced Interconversion of Protein Quaternary Structure Isoforms". Journal of Biological Chemistry. 280 (16): 15786–93. doi: 10.1074/jbc.M500218200 . PMID   15710608.
  10. Jaffe, Eileen K.; Lawrence, Sarah H. (2012). "Allostery and the dynamic oligomerization of porphobilinogen synthase". Archives of Biochemistry and Biophysics. 519 (2): 144–53. doi:10.1016/j.abb.2011.10.010. PMC   3291741 . PMID   22037356.
  11. 1 2 Lawrence, Sarah H.; Jaffe, Eileen K. (2008). "Expanding the concepts in protein structure-function relationships and enzyme kinetics: Teaching using morpheeins". Biochemistry and Molecular Biology Education. 36 (4): 274–283. doi:10.1002/bmb.20211. PMC   2575429 . PMID   19578473.
  12. 1 2 Monod, Jacques; Changeux, Jean-Pierre; Jacob, François (1963). "Allosteric proteins and cellular control systems". Journal of Molecular Biology. 6 (4): 306–29. doi:10.1016/S0022-2836(63)80091-1. PMID   13936070.
  13. 1 2 Monod, Jacques; Wyman, Jeffries; Changeux, Jean-Pierre (1965). "On the nature of allosteric transitions: A plausible model". Journal of Molecular Biology. 12: 88–118. doi:10.1016/S0022-2836(65)80285-6. PMID   14343300.
  14. Koshland, D.E. (1970). "7 The Molecular Basis for Enzyme Regulation". The Enzymes Volume 1. Vol. 1. pp. 341–396. doi:10.1016/S1874-6047(08)60170-5. ISBN   978-0-12-122701-2.
  15. Koshland, D. E.; Nemethy, G.; Filmer, D. (1966). "Comparison of Experimental Binding Data and Theoretical Models in Proteins Containing Subunits". Biochemistry. 5 (1): 365–85. doi:10.1021/bi00865a047. PMID   5938952.
  16. Frieden, C; Colman, R F (1967). "Glutamate dehydrogenase concentration as a determinant in the effect of purine nucleotides on the enzymatiuc activity". J. Biol. Chem. 242: 1705–1715. doi: 10.1016/S0021-9258(18)96059-X .
  17. Kurganov, B I (1982). Allosteric Enzymes: Kinetic Behaviour. Chichester: Wiley–Interscience. pp. 151–248. ISBN   978-0471101956.
  18. Friedrich, P (1984). Supramolecular Enzyme Organization: Quaternary Structure and Beyond. Oxford: Pergamon Press. pp. 66–71. ISBN   0-08-026376-3.
  19. Gerstein, Mark; Echols, Nathaniel (2004). "Exploring the range of protein flexibility, from a structural proteomics perspective". Current Opinion in Chemical Biology. 8 (1): 14–9. doi:10.1016/j.cbpa.2003.12.006. PMID   15036151.
  20. Carrell, Robin W; Lomas, David A (1997). "Conformational disease". The Lancet. 350 (9071): 134–8. doi:10.1016/S0140-6736(97)02073-4. PMID   9228977. S2CID   39124185.
  21. 1 2 Boone, A.N.; Brownsey, R.W.; Elliott, J.E.; Kulpa, J.E.; Lee, W.M. (2006). "Regulation of acetyl-CoA carboxylase". Biochemical Society Transactions. 34 (2): 223–7. doi:10.1042/BST20060223. PMID   16545081.
  22. Shen, Yang; Volrath, Sandra L.; Weatherly, Stephanie C.; Elich, Tedd D.; Tong, Liang (2004). "A Mechanism for the Potent Inhibition of Eukaryotic Acetyl-Coenzyme a Carboxylase by Soraphen A, a Macrocyclic Polyketide Natural Product". Molecular Cell. 16 (6): 881–91. doi: 10.1016/j.molcel.2004.11.034 . PMID   15610732.
  23. 1 2 3 Weissmann, Bernard; Wang, Ching-Te (1971). "Association-dissociation and abnormal kinetics of bovine .alpha.-acetylgalactosaminidase". Biochemistry. 10 (6): 1067–72. doi:10.1021/bi00782a021. PMID   5550813.
  24. 1 2 3 Weissmann, Bernard; Hinrichsen, Dorotea F. (1969). "Mammalian α-acetylgalactosaminidase. Occurrence, partial purification, and action on linkages in submaxillary mucins". Biochemistry. 8 (5): 2034–43. doi:10.1021/bi00833a038. PMID   5785223.
  25. De Zoysa Ariyananda, Lushanti; Colman, Roberta F. (2008). "Evaluation of Types of Interactions in Subunit Association in Bacillus subtilis Adenylosuccinate Lyase". Biochemistry. 47 (9): 2923–34. doi:10.1021/bi701400c. PMID   18237141.
  26. 1 2 3 Palenchar, Jennifer Brosius; Colman, Roberta F. (2003). "Characterization of a Mutant Bacillus subtilis Adenylosuccinate Lyase Equivalent to a Mutant Enzyme Found in Human Adenylosuccinate Lyase Deficiency: Asparagine 276 Plays an Important Structural Role". Biochemistry. 42 (7): 1831–41. doi:10.1021/bi020640+. PMID   12590570.
  27. Hohn, Thomas M.; Plattner, Ronald D. (1989). "Purification and characterization of the sesquiterpene cyclase aristolochene synthase from Penicillium roqueforti". Archives of Biochemistry and Biophysics. 272 (1): 137–43. doi:10.1016/0003-9861(89)90204-X. PMID   2544140.
  28. Caruthers, J. M.; Kang, I; Rynkiewicz, MJ; Cane, DE; Christianson, DW (2000). "Crystal Structure Determination of Aristolochene Synthase from the Blue Cheese Mold, Penicillium roqueforti". Journal of Biological Chemistry. 275 (33): 25533–9. doi: 10.1074/jbc.M000433200 . PMID   10825154.
  29. Jerebzoff-Quintin, Simonne; Jerebzoff, Stephan (1985). "L-Asparaginase activity in Leptosphaeria michotii. Isolation and properties of two forms of the enzyme". Physiologia Plantarum. 64: 74–80. doi:10.1111/j.1399-3054.1985.tb01215.x.
  30. Yun, Mi-Kyung; Nourse, Amanda; White, Stephen W.; Rock, Charles O.; Heath, Richard J. (2007). "Crystal Structure and Allosteric Regulation of the Cytoplasmic Escherichia coli l-Asparaginase I". Journal of Molecular Biology. 369 (3): 794–811. doi:10.1016/j.jmb.2007.03.061. PMC   1991333 . PMID   17451745.
  31. Garel, J.-R. (1980). "Sequential Folding of a Bifunctional Allosteric Protein". Proceedings of the National Academy of Sciences. 77 (6): 3379–3383. Bibcode:1980PNAS...77.3379G. doi: 10.1073/pnas.77.6.3379 . JSTOR   8892. PMC   349619 . PMID   6774337.
  32. 1 2 Kotaka, M.; Ren, J.; Lockyer, M.; Hawkins, A. R.; Stammers, D. K. (2006). "Structures of R- and T-state Escherichia coli Aspartokinase III: MECHANISMS OF THE ALLOSTERIC TRANSITION AND INHIBITION BY LYSINE". Journal of Biological Chemistry. 281 (42): 31544–52. doi: 10.1074/jbc.M605886200 . PMID   16905770.
  33. Ogilvie, JW; Vickers, LP; Clark, RB; Jones, MM (1975). "Aspartokinase I-homoserine dehydrogenase I of Escherichia coli K12 (lambda). Activation by monovalent cations and an analysis of the effect of the adenosine triphosphate-magnesium ion complex on this activation process". The Journal of Biological Chemistry. 250 (4): 1242–50. doi: 10.1016/S0021-9258(19)41805-X . PMID   163250.
  34. 1 2 Trompier, D.; Alibert, M; Davanture, S; Hamon, Y; Pierres, M; Chimini, G (2006). "Transition from Dimers to Higher Oligomeric Forms Occurs during the ATPase Cycle of the ABCA1 Transporter". Journal of Biological Chemistry. 281 (29): 20283–90. doi: 10.1074/jbc.M601072200 . PMID   16709568.
  35. 1 2 Eisenstein, Edward; Beckett, Dorothy (1999). "Dimerization of theEscherichiacoliBiotin Repressor: Corepressor Function in Protein Assembly". Biochemistry. 38 (40): 13077–84. doi:10.1021/bi991241q. PMID   10529178.
  36. Streaker, Emily D.; Beckett, Dorothy (1998). "Coupling of Site-Specific DNA Binding to Protein Dimerization in Assembly of the Biotin Repressor−Biotin Operator Complex". Biochemistry. 37 (9): 3210–9. doi:10.1021/bi9715019. PMID   9485476.
  37. Vamvaca, Katherina; Butz, Maren; Walter, Kai U.; Taylor, Sean V.; Hilvert, Donald (2005). "Simultaneous optimization of enzyme activity and quaternary structure by directed evolution". Protein Science. 14 (8): 2103–14. doi:10.1110/ps.051431605. PMC   2279322 . PMID   15987889.
  38. 1 2 3 4 5 Tong, E. K.; Duckworth, Harry W. (1975). "Quaternary structure of citrate synthase from Escherichia coli K 12". Biochemistry. 14 (2): 235–41. doi:10.1021/bi00673a007. PMID   1091285.
  39. Bewley, Carole A.; Gustafson, Kirk R.; Boyd, Michael R.; Covell, David G.; Bax, Ad; Clore, G. Marius; Gronenborn, Angela M. (1998). "Solution structure of cyanovirin-N, a potent HIV-inactivating protein". Nature Structural Biology. 5 (7): 571–8. doi:10.1038/828. PMID   9665171. S2CID   11367037.
  40. Yang, Fan; Bewley, Carole A; Louis, John M; Gustafson, Kirk R; Boyd, Michael R; Gronenborn, Angela M; Clore, G.Marius; Wlodawer, Alexander (1999). "Crystal structure of cyanovirin-N, a potent HIV-inactivating protein, shows unexpected domain swapping". Journal of Molecular Biology. 288 (3): 403–12. doi:10.1006/jmbi.1999.2693. PMID   10329150. S2CID   308708.
  41. 1 2 Barrientos, LG; Gronenborn, AM (2005). "The highly specific carbohydrate-binding protein cyanovirin-N: Structure, anti-HIV/Ebola activity and possibilities for therapy". Mini Reviews in Medicinal Chemistry. 5 (1): 21–31. doi:10.2174/1389557053402783. PMID   15638789.
  42. 1 2 Barrientos, LG; Louis, JM; Botos, I; Mori, T; Han, Z; O'Keefe, BR; Boyd, MR; Wlodawer, A; et al. (2002). "The domain-swapped dimer of cyanovirin-N is in a metastable folded state: Reconciliation of X-ray and NMR structures". Structure. 10 (5): 673–86. doi: 10.1016/S0969-2126(02)00758-X . PMID   12015150.
  43. 1 2 3 Rochet, Jean-Christophe; Brownie, Edward R.; Oikawa, Kim; Hicks, Leslie D.; Fraser, Marie E.; James, Michael N. G.; Kay, Cyril M.; Bridger, William A.; et al. (2000). "Pig Heart CoA Transferase Exists as Two Oligomeric Forms Separated by a Large Kinetic Barrier". Biochemistry. 39 (37): 11291–302. doi:10.1021/bi0003184. PMID   10985774.
  44. Frank, Nina; Kery, Vladimir; MacLean, Kenneth N.; Kraus, Jan P. (2006). "Solvent-Accessible Cysteines in Human Cystathionine β-Synthase: Crucial Role of Cysteine 431 inS-Adenosyl-l-methionine Binding". Biochemistry. 45 (36): 11021–9. doi:10.1021/bi060737m. PMID   16953589.
  45. 1 2 Sen, Suvajit; Banerjee, Ruma (2007). "A Pathogenic Linked Mutation in the Catalytic Core of Human Cystathionine β-Synthase Disrupts Allosteric Regulation and Allows Kinetic Characterization of a Full-Length Dimer". Biochemistry. 46 (13): 4110–6. doi:10.1021/bi602617f. PMC   3204387 . PMID   17352495.
  46. Kery, Vladimir; Poneleit, Loelle; Kraus, Jan P. (1998). "Trypsin Cleavage of Human Cystathionine β-Synthase into an Evolutionarily Conserved Active Core: Structural and Functional Consequences". Archives of Biochemistry and Biophysics. 355 (2): 222–32. doi:10.1006/abbi.1998.0723. PMID   9675031.
  47. Shan, Xiaoyin; Kruger, Warren D. (1998). "Correction of disease-causing CBS mutations in yeast". Nature Genetics. 19 (1): 91–3. doi:10.1038/ng0598-91. PMID   9590298. S2CID   47102642.
  48. 1 2 Antonini, E; Brunori, M; Bruzzesi, R; Chiancone, E; Massey, V (1966). "Association-dissociation phenomena of D-amino acid oxidase". The Journal of Biological Chemistry. 241 (10): 2358–66. doi: 10.1016/S0021-9258(18)96629-9 . PMID   4380380.
  49. 1 2 Massey, V; Curti, B; Ganther, H (1966). "A temperature-dependent conformational change in D-amino acid oxidase and its effect on catalysis". The Journal of Biological Chemistry. 241 (10): 2347–57. doi: 10.1016/S0021-9258(18)96628-7 . PMID   5911617.
  50. 1 2 3 4 Babady, N. E.; Pang, Y.-P.; Elpeleg, O.; Isaya, G. (2007). "Cryptic proteolytic activity of dihydrolipoamide dehydrogenase". Proceedings of the National Academy of Sciences. 104 (15): 6158–63. Bibcode:2007PNAS..104.6158B. doi: 10.1073/pnas.0610618104 . PMC   1851069 . PMID   17404228.
  51. Muiswinkel-Voetberg, H.; Visser, Jaap; Veeger, Cornelis (1973). "Conformational Studies on Lipoamide Dehydrogenase from Pig Heart. 1. Interconversion of Dissociable and Non-Dissociable Forms". European Journal of Biochemistry. 33 (2): 265–70. doi:10.1111/j.1432-1033.1973.tb02679.x. PMID   4348439.
  52. Klyachko, N. L.; Shchedrina, VA; Efimov, AV; Kazakov, SV; Gazaryan, IG; Kristal, BS; Brown, AM (2005). "PH-dependent Substrate Preference of Pig Heart Lipoamide Dehydrogenase Varies with Oligomeric State: RESPONSE TO MITOCHONDRIAL MATRIX ACIDIFICATION". Journal of Biological Chemistry. 280 (16): 16106–14. doi: 10.1074/jbc.M414285200 . PMID   15710613.
  53. Muiswinkel-Voetberg, H.; Veeger, Cornelis (1973). "Conformational Studies on Lipoamide Dehydrogenase from Pig Heart. 2. Spectroscopic Studies on the Apoenzyme and the Monomeric and Dimeric Forms". European Journal of Biochemistry. 33 (2): 271–8. doi: 10.1111/j.1432-1033.1973.tb02680.x . PMID   4348440.
  54. 1 2 3 4 Saxena, Ashima; Hensley, Preston; Osborne, James C.; Fleming, Patrick J. (1985). "The pH-dependent Subunit Dissociation and Catalytic Activity of Bovine Dopamine β-Hydroxylase". Journal of Biological Chemistry. 260 (6): 3386–92. doi: 10.1016/S0021-9258(19)83633-5 . PMID   3972830.
  55. 1 2 3 4 Dhawan, S; Hensley, P; Osborne Jr, JC; Fleming, PJ (1986). "Adenosine 5'-diphosphate-dependent subunit dissociation of bovine dopamine beta-hydroxylase". The Journal of Biological Chemistry. 261 (17): 7680–4. doi: 10.1016/S0021-9258(19)57453-1 . PMID   3711102.
  56. 1 2 3 4 Stewart, L C; Klinman, J P (1988). "Dopamine Beta-Hydroxylase of Adrenal Chromaffin Granules: Structure and Function". Annual Review of Biochemistry. 57: 551–92. doi:10.1146/annurev.bi.57.070188.003003. PMID   3052283.
  57. Kuzuguchi, T.; Morita, Y; Sagami, I; Sagami, H; Ogura, K (1999). "Human Geranylgeranyl Diphosphate Synthase. CDNA CLONING AND EXPRESSION". Journal of Biological Chemistry. 274 (9): 5888–94. doi: 10.1074/jbc.274.9.5888 . PMID   10026212.
  58. 1 2 Kavanagh, K. L.; Dunford, JE; Bunkoczi, G; Russell, RG; Oppermann, U (2006). "The Crystal Structure of Human Geranylgeranyl Pyrophosphate Synthase Reveals a Novel Hexameric Arrangement and Inhibitory Product Binding" (PDF). Journal of Biological Chemistry. 281 (31): 22004–12. doi: 10.1074/jbc.M602603200 . PMID   16698791.
  59. Miyagi, Y.; Matsumura, Y.; Sagami, H. (2007). "Human Geranylgeranyl Diphosphate Synthase is an Octamer in Solution". Journal of Biochemistry. 142 (3): 377–81. doi:10.1093/jb/mvm144. PMID   17646172.
  60. Snook, Christopher F.; Tipton, Peter A.; Beamer, Lesa J. (2003). "Crystal Structure of GDP-Mannose Dehydrogenase: A Key Enzyme of Alginate Biosynthesis inP. Aeruginosa". Biochemistry. 42 (16): 4658–68. doi:10.1021/bi027328k. PMID   12705829.
  61. Roychoudhury, S; May, TB; Gill, JF; Singh, SK; Feingold, DS; Chakrabarty, AM (1989). "Purification and characterization of guanosine diphospho-D-mannose dehydrogenase. A key enzyme in the biosynthesis of alginate by Pseudomonas aeruginosa". The Journal of Biological Chemistry. 264 (16): 9380–5. doi: 10.1016/S0021-9258(18)60542-3 . PMID   2470755.
  62. 1 2 Naught, Laura E.; Gilbert, Sunny; Imhoff, Rebecca; Snook, Christopher; Beamer, Lesa; Tipton, Peter (2002). "Allosterism and Cooperativity inPseudomonas aeruginosaGDP-Mannose Dehydrogenase". Biochemistry. 41 (30): 9637–45. doi:10.1021/bi025862m. PMID   12135385.
  63. 1 2 Fisher, Harvey F. (2006). "Glutamate Dehydrogenase—ligand Complexes and Their Relationship to the Mechanism of the Reaction". Advances in Enzymology and Related Areas of Molecular Biology. Advances in Enzymology - and Related Areas of Molecular Biology. Vol. 39. pp.  369–417. doi:10.1002/9780470122846.ch6. ISBN   978-0-470-12284-6. PMID   4147773.
  64. Huang, CY; Frieden, C (1972). "The mechanism of ligand-induced structural changes in glutamate dehydrogenase. Studies of the rate of depolymerization and isomerization effected by coenzymes and guanine nucleotides". The Journal of Biological Chemistry. 247 (11): 3638–46. doi: 10.1016/S0021-9258(19)45188-0 . PMID   4402280.
  65. 1 2 Kim, Sang Suk; Choi, I.-G.; Kim, Sung-Hou; Yu, Y. G. (1999). "Molecular cloning, expression, and characterization of a thermostable glutamate racemase from a hyperthermophilic bacterium, Aquifex pyrophilus". Extremophiles. 3 (3): 175–83. doi:10.1007/s007920050114. PMID   10484173. S2CID   709039.
  66. 1 2 Lundqvist, Tomas; Fisher, Stewart L.; Kern, Gunther; Folmer, Rutger H. A.; Xue, Yafeng; Newton, D. Trevor; Keating, Thomas A.; Alm, Richard A.; et al. (2007). "Exploitation of structural and regulatory diversity in glutamate racemases". Nature. 447 (7146): 817–22. Bibcode:2007Natur.447..817L. doi:10.1038/nature05689. PMID   17568739. S2CID   4408683.
  67. 1 2 May, Melissa; Mehboob, Shahila; Mulhearn, Debbie C.; Wang, Zhiqiang; Yu, Huidong; Thatcher, Gregory R.J.; Santarsiero, Bernard D.; Johnson, Michael E.; et al. (2007). "Structural and Functional Analysis of Two Glutamate Racemase Isozymes from Bacillus anthracis and Implications for Inhibitor Design". Journal of Molecular Biology. 371 (5): 1219–37. doi:10.1016/j.jmb.2007.05.093. PMC   2736553 . PMID   17610893.
  68. 1 2 Taal, Makie A.; Sedelnikova, Svetlana E.; Ruzheinikov, Sergey N.; Baker, Patrick J.; Rice, David W. (2004). "Expression, purification and preliminary X-ray analysis of crystals ofBacillus subtilisglutamate racemase". Acta Crystallographica Section D. 60 (11): 2031–4. Bibcode:2004AcCrD..60.2031T. doi: 10.1107/S0907444904021134 . PMID   15502318.
  69. 1 2 Kim, Kook-Han; Bong, Young-Jong; Park, Joon Kyu; Shin, Key-Jung; Hwang, Kwang Yeon; Kim, Eunice Eunkyeong (2007). "Structural Basis for Glutamate Racemase Inhibition". Journal of Molecular Biology. 372 (2): 434–43. doi:10.1016/j.jmb.2007.05.003. PMID   17658548.
  70. Ashiuchi, M.; Kuwana, E; Yamamoto, T; Komatsu, K; Soda, K; Misono, H (2002). "Glutamate Racemase is an Endogenous DNA Gyrase Inhibitor". Journal of Biological Chemistry. 277 (42): 39070–3. doi: 10.1074/jbc.C200253200 . hdl: 10126/3383 . PMID   12213801.
  71. Ashiuchi, M.; Tani, K.; Soda, K.; Misono, H. (1998). "Properties of Glutamate Racemase from Bacillus subtilis IFO 3336 Producing Poly- -Glutamate". Journal of Biochemistry. 123 (6): 1156–63. doi:10.1093/oxfordjournals.jbchem.a022055. PMID   9604005.
  72. Sengupta, S.; Ghosh, S.; Nagaraja, V. (2008). "Moonlighting function of glutamate racemase from Mycobacterium tuberculosis: Racemization and DNA gyrase inhibition are two independent activities of the enzyme". Microbiology. 154 (9): 2796–803. doi: 10.1099/mic.0.2008/020933-0 . PMID   18757813.
  73. Sirover, Michael A (1999). "New insights into an old protein: The functional diversity of mammalian glyceraldehyde-3-phosphate dehydrogenase". Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology. 1432 (2): 159–84. doi:10.1016/S0167-4838(99)00119-3. PMID   10407139.
  74. Constantinides, SM; Deal Jr, WC (1969). "Reversible dissociation of tetrameric rabbit muscle glyceraldehyde 3-phosphate dehydrogenase into dimers or monomers by adenosine triphosphate". The Journal of Biological Chemistry. 244 (20): 5695–702. doi: 10.1016/S0021-9258(18)63615-4 . PMID   4312250.
  75. Kumagai, H; Sakai, H (1983). "A porcine brain protein (35 K protein) which bundles microtubules and its identification as glyceraldehyde 3-phosphate dehydrogenase". Journal of Biochemistry. 93 (5): 1259–69. doi:10.1093/oxfordjournals.jbchem.a134260. PMID   6885722.
  76. 1 2 De Riel, Jon K.; Paulus, Henry (1978). "Subunit dissociation in the allosteric regulation of glycerol kinase from Escherichia coli. 2. Physical evidence". Biochemistry. 17 (24): 5141–6. doi:10.1021/bi00617a011. PMID   215195.
  77. 1 2 De Riel, Jon K.; Paulus, Henry (1978). "Subunit dissociation in the allosteric regulation of glycerol kinase from Escherichia coli. 1. Kinetic evidence". Biochemistry. 17 (24): 5134–40. doi:10.1021/bi00617a010. PMID   215194.
  78. 1 2 De Riel, Jon K.; Paulus, Henry (1978). "Subunit dissociation in the allosteric regulation of glycerol kinase from Escherichia coli. 3. Role in desensitization". Biochemistry. 17 (24): 5146–50. doi:10.1021/bi00617a012. PMID   31903.
  79. 1 2 Feese, Michael D; Faber, H Rick; Bystrom, Cory E; Pettigrew, Donald W; Remington, S James (1998). "Glycerol kinase from Escherichia coli and an Ala65→Thr mutant: The crystal structures reveal conformational changes with implications for allosteric regulation". Structure. 6 (11): 1407–18. doi: 10.1016/S0969-2126(98)00140-3 . PMID   9817843.
  80. 1 2 Bystrom, Cory E.; Pettigrew, Donald W.; Branchaud, Bruce P.; O'Brien, Patrick; Remington, S. James (1999). "Crystal Structures ofEscherichia coliGlycerol Kinase Variant S58→W in Complex with Nonhydrolyzable ATP Analogues Reveal a Putative Active Conformation of the Enzyme as a Result of Domain Motion". Biochemistry. 38 (12): 3508–18. doi:10.1021/bi982460z. PMID   10090737.
  81. 1 2 Deprez, Eric; Tauc, Patrick; Leh, Hervé; Mouscadet, Jean-François; Auclair, Christian; Brochon, Jean-Claude (2000). "Oligomeric States of the HIV-1 Integrase As Measured by Time-Resolved Fluorescence Anisotropy". Biochemistry. 39 (31): 9275–84. doi:10.1021/bi000397j. PMID   10924120.
  82. 1 2 Deprez, E.; Tauc, P.; Leh, H.; Mouscadet, J.-F.; Auclair, C.; Hawkins, M. E.; Brochon, J.-C. (2001). "DNA binding induces dissociation of the multimeric form of HIV-1 integrase: A time-resolved fluorescence anisotropy study". Proceedings of the National Academy of Sciences. 98 (18): 10090–5. Bibcode:2001PNAS...9810090D. doi: 10.1073/pnas.181024498 . PMC   56920 . PMID   11504911.
  83. 1 2 3 Faure, A. l.; Calmels, C; Desjobert, C; Castroviejo, M; Caumont-Sarcos, A; Tarrago-Litvak, L; Litvak, S; Parissi, V (2005). "HIV-1 integrase crosslinked oligomers are active in vitro". Nucleic Acids Research. 33 (3): 977–86. doi:10.1093/nar/gki241. PMC   549407 . PMID   15718297.
  84. 1 2 Guiot, E.; Carayon, K; Delelis, O; Simon, F; Tauc, P; Zubin, E; Gottikh, M; Mouscadet, JF; et al. (2006). "Relationship between the Oligomeric Status of HIV-1 Integrase on DNA and Enzymatic Activity". Journal of Biological Chemistry. 281 (32): 22707–19. doi: 10.1074/jbc.M602198200 . PMID   16774912.
  85. Fieulaine, S.; Morera, S; Poncet, S; Monedero, V; Gueguen-Chaignon, V; Galinier, A; Janin, J; Deutscher, J; et al. (2001). "X-ray structure of HPr kinase: A bacterial protein kinase with a P-loop nucleotide-binding domain". The EMBO Journal. 20 (15): 3917–27. doi:10.1093/emboj/20.15.3917. PMC   149164 . PMID   11483495.
  86. Márquez, José Antonio; Hasenbein, Sonja; Koch, Brigitte; Fieulaine, Sonia; Nessler, Sylvie; Russell, Robert B.; Hengstenberg, Wolfgang; Scheffzek, Klaus (2002). "Structure of the full-length HPr kinase/phosphatase from Staphylococcus xylosus at 1.95 Å resolution: Mimicking the product/substrate of the phospho transfer reactions". Proceedings of the National Academy of Sciences. 99 (6): 3458–63. Bibcode:2002PNAS...99.3458M. doi: 10.1073/pnas.052461499 . JSTOR   3058148. PMC   122545 . PMID   11904409.
  87. Allen, Gregory S.; Steinhauer, Katrin; Hillen, Wolfgang; Stülke, Jörg; Brennan, Richard G. (2003). "Crystal Structure of HPr Kinase/Phosphatase from Mycoplasma pneumoniae". Journal of Molecular Biology. 326 (4): 1203–17. doi:10.1016/S0022-2836(02)01378-5. PMID   12589763.
  88. Poncet, Sandrine; Mijakovic, Ivan; Nessler, Sylvie; Gueguen-Chaignon, Virginie; Chaptal, Vincent; Galinier, Anne; Boël, Grégory; Mazé, Alain; et al. (2004). "HPr kinase/phosphorylase, a Walker motif A-containing bifunctional sensor enzyme controlling catabolite repression in Gram-positive bacteria". Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics. 1697 (1–2): 123–35. doi:10.1016/j.bbapap.2003.11.018. PMID   15023355.
  89. 1 2 3 4 5 Ramstrom, H.; Sanglier, S; Leize-Wagner, E; Philippe, C; Van Dorsselaer, A; Haiech, J (2002). "Properties and Regulation of the Bifunctional Enzyme HPr Kinase/Phosphatase in Bacillus subtilis". Journal of Biological Chemistry. 278 (2): 1174–85. doi: 10.1074/jbc.M209052200 . PMID   12411438.
  90. Jault, J.-M.; Fieulaine, S; Nessler, S; Gonzalo, P; Di Pietro, A; Deutscher, J; Galinier, A (2000). "The HPr Kinase from Bacillus subtilis is a Homo-oligomeric Enzyme Which Exhibits Strong Positive Cooperativity for Nucleotide and Fructose 1,6-Bisphosphate Binding" (PDF). Journal of Biological Chemistry. 275 (3): 1773–80. doi: 10.1074/jbc.275.3.1773 . PMID   10636874.
  91. Clarke, Anthony R.; Waldman, Adam D.B.; Munro, Ian; Holbrook, J.John (1985). "Changes in the state of subunit association of lactate dehydrogenase from Bacillus stearothermophilus". Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology. 828 (3): 375–9. doi:10.1016/0167-4838(85)90319-X. PMID   3986214.
  92. 1 2 3 4 5 Clarke, Anthony R.; Waldman, Adam D.B.; Hart, Keith W.; John Holbrook, J. (1985). "The rates of defined changes in protein structure during the catalytic cycle of lactate dehydrogenase". Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology. 829 (3): 397–407. doi:10.1016/0167-4838(85)90250-X. PMID   4005269.
  93. Clarke, Anthony R.; Wigley, Dale B.; Barstow, David A.; Chia, William N.; Atkinson, Tony; Holbrook, J.John (1987). "A single amino acid substitution deregulates a bacterial lactate dehydrogenase and stabilizes its tetrameric structure". Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology. 913 (1): 72–80. doi:10.1016/0167-4838(87)90234-2. PMID   3580377.
  94. Cameron, Alexander D.; Roper, David I.; Moreton, Kathleen M.; Muirhead, Hilary; Holbrook, J.John; Wigley, Dale B. (1994). "Allosteric Activation in Bacillus stearothermophilus Lactate Dehydrogenase Investigated by an X-ray Crystallographic Analysis of a Mutant Designed to Prevent Tetramerization of the Enzyme". Journal of Molecular Biology. 238 (4): 615–25. doi:10.1006/jmbi.1994.1318. PMID   8176749.
  95. 1 2 3 Roudiak, Stanislav G.; Shrader, Thomas E. (1998). "Functional Role of the N-Terminal Region of the Lon Protease fromMycobacterium smegmatis". Biochemistry. 37 (32): 11255–63. doi:10.1021/bi980945h. PMID   9698372.
  96. 1 2 3 Rudyak, Stanislav G.; Brenowitz, Michael; Shrader, Thomas E. (2001). "Mg2+-Linked Oligomerization Modulates the Catalytic Activity of the Lon (La) Protease from Mycobacterium smegmatis". Biochemistry. 40 (31): 9317–23. doi:10.1021/bi0102508. PMID   11478899.
  97. 1 2 Vineyard, Diana; Patterson-Ward, Jessica; Lee, Irene (2006). "Single-Turnover Kinetic Experiments Confirm the Existence of High- and Low-Affinity ATPase Sites inEscherichia coliLon Protease". Biochemistry. 45 (14): 4602–10. doi:10.1021/bi052377t. PMC   2515378 . PMID   16584195.
  98. 1 2 Yang, Zhiru; Lanks, Charles W.; Tong, Liang (2002). "Molecular Mechanism for the Regulation of Human Mitochondrial NAD(P)+-Dependent Malic Enzyme by ATP and Fumarate". Structure. 10 (7): 951–60. doi: 10.1016/S0969-2126(02)00788-8 . PMID   12121650.
  99. 1 2 Gerald e, Edwards; Carlos s, Andreo (1992). "NADP-malic enzyme from plants". Phytochemistry. 31 (6): 1845–57. Bibcode:1992PChem..31.1845G. doi:10.1016/0031-9422(92)80322-6. PMID   1368216.
  100. Hsieh, J.-Y.; Chen, S.-H.; Hung, H.-C. (2009). "Functional Roles of the Tetramer Organization of Malic Enzyme". Journal of Biological Chemistry. 284 (27): 18096–105. doi: 10.1074/jbc.M109.005082 . PMC   2709377 . PMID   19416979.
  101. Poole, Leslie B. (2005). "Bacterial defenses against oxidants: Mechanistic features of cysteine-based peroxidases and their flavoprotein reductases". Archives of Biochemistry and Biophysics. 433 (1): 240–54. doi:10.1016/j.abb.2004.09.006. PMID   15581580.
  102. Aran, Martin; Ferrero, Diego S.; Pagano, Eduardo; Wolosiuk, Ricardo A. (2009). "Typical 2-Cys peroxiredoxins - modulation by covalent transformations and noncovalent interactions". FEBS Journal. 276 (9): 2478–93. doi:10.1111/j.1742-4658.2009.06984.x. hdl: 11336/20656 . PMID   19476489. S2CID   1698327.
  103. Bjørgo, Elisa; De Carvalho, Raquel Margarida Negrão; Flatmark, Torgeir (2001). "A comparison of kinetic and regulatory properties of the tetrameric and dimeric forms of wild-type and Thr427→Pro mutant human phenylalanine hydroxylase". European Journal of Biochemistry. 268 (4): 997–1005. doi:10.1046/j.1432-1327.2001.01958.x. PMID   11179966.
  104. Martinez, Aurora; Knappskog, Per M.; Olafsdottir, Sigridur; Døskeland, Anne P.; Eiken, Hans Geir; Svebak, Randi Myrseth; Bozzini, MeriLisa; Apold, Jaran; et al. (1995). "Expression of recombinant human phenylalanine hydroxylase as fusion protein in Escherichia coli circumvents proteolytic degradation by host cell proteases. Isolation and characterization of the wild-type enzyme". The Biochemical Journal. 306 (2): 589–97. doi:10.1042/bj3060589. PMC   1136558 . PMID   7887915.
  105. Knappskog, Per M.; Flatmark, Torgeir; Aarden, Johanna M.; Haavik, Jan; Martinez, Aurora (1996). "Structure/Function Relationships in Human Phenylalanine Hydroxylase. Effect of Terminal Deletions on the Oligomerization, Activation and Cooperativity of Substrate Binding to the Enzyme". European Journal of Biochemistry. 242 (3): 813–21. doi: 10.1111/j.1432-1033.1996.0813r.x . PMID   9022714.
  106. Phillips, Robert S.; Parniak, Michael A.; Kaufman, Seymour (1984). "Spectroscopic investigation of ligand interaction with hepatic phenylalanine hydroxylase: Evidence for a conformational change associated with activation". Biochemistry. 23 (17): 3836–42. doi:10.1021/bi00312a007. PMID   6487579.
  107. Fusetti, F.; Erlandsen, H; Flatmark, T; Stevens, RC (1998). "Structure of Tetrameric Human Phenylalanine Hydroxylase and Its Implications for Phenylketonuria". Journal of Biological Chemistry. 273 (27): 16962–7. doi: 10.1074/jbc.273.27.16962 . PMID   9642259.
  108. 1 2 3 4 5 6 Wohl, RC; Markus, G (1972). "Phosphoenolpyruvate carboxylase of Escherichia coli. Purification and some properties". The Journal of Biological Chemistry. 247 (18): 5785–92. doi: 10.1016/S0021-9258(19)44827-8 . PMID   4560418.
  109. Kai, Yasushi; Matsumura, Hiroyoshi; Izui, Katsura (2003). "Phosphoenolpyruvate carboxylase: Three-dimensional structure and molecular mechanisms". Archives of Biochemistry and Biophysics. 414 (2): 170–9. doi:10.1016/S0003-9861(03)00170-X. PMID   12781768.
  110. 1 2 3 Xu, Jing; Oshima, Tairo; Yoshida, Masasuke (1990). "Tetramer-dimer conversion of phosphofructokinase from Thermus thermophilus induced by its allosteric effectors". Journal of Molecular Biology. 215 (4): 597–606. doi:10.1016/S0022-2836(05)80171-8. PMID   2146397.
  111. Jolley Jr, RL; Mason, HS (1965). "The Multiple Forms of Mushroom Tyrosinase. Interconversion". The Journal of Biological Chemistry. 240: PC1489–91. doi: 10.1016/S0021-9258(18)97603-9 . PMID   14284774.
  112. Jolley Jr, RL; Robb, DA; Mason, HS (1969). "The multiple forms of mushroom tyrosinase. Association-dissociation phenomena". The Journal of Biological Chemistry. 244 (6): 1593–9. doi: 10.1016/S0021-9258(18)91800-4 . PMID   4975157.
  113. Mallette, MF; Dawson, CR (1949). "On the nature of highly purified mushroom tyrosinase preparations". Archives of Biochemistry. 23 (1): 29–44. PMID   18135760.
  114. 1 2 Chazarra, Soledad; García-Carmona, Francisco; Cabanes, Juana (2001). "Hysteresis and Positive Cooperativity of Iceberg Lettuce Polyphenol Oxidase". Biochemical and Biophysical Research Communications. 289 (3): 769–75. doi:10.1006/bbrc.2001.6014. PMID   11726215.
  115. Harel, E.; Mayer, A.M. (1968). "Interconversion of sub-units of catechol oxidase from apple chloroplasts". Phytochemistry. 7 (2): 199–204. Bibcode:1968PChem...7..199H. doi:10.1016/S0031-9422(00)86315-3.
  116. 1 2 Jaffe EK, Lawrence SH (March 2012). "Allostery and the dynamic oligomerization of porphobilinogen synthase". Arch. Biochem. Biophys. 519 (2): 144–53. doi:10.1016/j.abb.2011.10.010. PMC   3291741 . PMID   22037356.
  117. Breinig S, Kervinen J, Stith L, Wasson AS, Fairman R, Wlodawer A, Zdanov A, Jaffe EK (September 2003). "Control of tetrapyrrole biosynthesis by alternate quaternary forms of porphobilinogen synthase". Nat. Struct. Biol. 10 (9): 757–63. doi:10.1038/nsb963. PMID   12897770. S2CID   24188785.
  118. 1 2 Schulz, Ju¨Rgen; Sparmann, Gisela; Hofmann, Eberhard (1975). "Alanine-mediated reversible inactivation of tumour pyruvate kinase caused by a tetramer-dimer transition". FEBS Letters. 50 (3): 346–50. doi: 10.1016/0014-5793(75)90064-2 . PMID   1116605. S2CID   5665440.
  119. 1 2 Ibsen, KH; Schiller, KW; Haas, TA (1971). "Interconvertible kinetic and physical forms of human erythrocyte pyruvate kinase". The Journal of Biological Chemistry. 246 (5): 1233–40. doi: 10.1016/S0021-9258(19)76963-4 . PMID   5545066.
  120. Liu, Yanshun; Gotte, Giovanni; Libonati, Massimo; Eisenberg, David (2009). "Structures of the two 3D domain-swapped RNase a trimers". Protein Science. 11 (2): 371–80. doi:10.1110/ps.36602. PMC   2373430 . PMID   11790847.
  121. 1 2 Gotte, Giovanni; Bertoldi, Mariarita; Libonati, Massimo (1999). "Structural versatility of bovine ribonuclease A. Distinct conformers of trimeric and tetrameric aggregates of the enzyme". European Journal of Biochemistry. 265 (2): 680–7. doi: 10.1046/j.1432-1327.1999.00761.x . PMID   10504400.
  122. Gotte, Giovanni; Laurents, Douglas V.; Libonati, Massimo (2006). "Three-dimensional domain-swapped oligomers of ribonuclease A: Identification of a fifth tetramer, pentamers and hexamers, and detection of trace heptameric, octameric and nonameric species". Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics. 1764 (1): 44–54. doi:10.1016/j.bbapap.2005.10.011. PMID   16310422.
  123. 1 2 Gotte, Giovanni; Libonati, Massimo (1998). "Two different forms of aggregated dimers of ribonuclease A". Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology. 1386 (1): 106–112. doi:10.1016/S0167-4838(98)00087-9. PMID   9675255.
  124. 1 2 Libonati, Massimo; Gotte, Giovanni (2004). "Oligomerization of bovine ribonuclease A: Structural and functional features of its multimers". Biochemical Journal. 380 (2): 311–27. doi:10.1042/BJ20031922. PMC   1224197 . PMID   15104538.
  125. 1 2 Libonati, M. (2004). "Biological actions of the oligomers of ribonuclease A". Cellular and Molecular Life Sciences. 61 (19–20): 2431–6. doi:10.1007/s00018-004-4302-x. PMID   15526151. S2CID   8769502.
  126. 1 2 Libonati, M; Bertoldi, M; Sorrentino, S (1996). "The activity on double-stranded RNA of aggregates of ribonuclease a higher than dimers increases as a function of the size of the aggregates". The Biochemical Journal. 318 (1): 287–90. doi:10.1042/bj3180287. PMC   1217620 . PMID   8761484.
  127. 1 2 Libonati, M.; Gotte, G.; Vottariello, F. (2008). "A Novel Biological Actions Acquired by Ribonuclease Through Oligomerization". Current Pharmaceutical Biotechnology. 9 (3): 200–9. doi:10.2174/138920108784567308. PMID   18673285.
  128. Kashlan, Ossama B.; Cooperman, Barry S. (2003). "Comprehensive Model for Allosteric Regulation of Mammalian Ribonucleotide Reductase: Refinements and Consequences†". Biochemistry. 42 (6): 1696–706. doi:10.1021/bi020634d. PMID   12578384.
  129. Kashlan, Ossama B.; Scott, Charles P.; Lear, James D.; Cooperman, Barry S. (2002). "A Comprehensive Model for the Allosteric Regulation of Mammalian Ribonucleotide Reductase. Functional Consequences of ATP- and dATP-Induced Oligomerization of the Large Subunit†". Biochemistry. 41 (2): 462–74. doi:10.1021/bi011653a. PMID   11781084.
  130. Eriksson, Mathias; Uhlin, Ulla; Ramaswamy, S; Ekberg, Monica; Regnström, Karin; Sjöberg, Britt-Marie; Eklund, Hans (1997). "Binding of allosteric effectors to ribonucleotide reductase protein R1: Reduction of active-site cysteines promotes substrate binding". Structure. 5 (8): 1077–92. doi: 10.1016/S0969-2126(97)00259-1 . PMID   9309223.
  131. 1 2 Fairman, James Wesley; Wijerathna, Sanath Ranjan; Ahmad, Md Faiz; Xu, Hai; Nakano, Ryo; Jha, Shalini; Prendergast, Jay; Welin, R Martin; et al. (2011). "Structural basis for allosteric regulation of human ribonucleotide reductase by nucleotide-induced oligomerization". Nature Structural & Molecular Biology. 18 (3): 316–22. doi:10.1038/nsmb.2007. PMC   3101628 . PMID   21336276.
  132. 1 2 Hohman, R.J.; Guitton, M.C.; Véron, M. (1984). "Purification of S-adenosyl-l-homocysteine hydrolase from Dictyostelium discoideum: Reversible inactivation by cAMP and 2′-deoxyadenosine". Archives of Biochemistry and Biophysics. 233 (2): 785–95. doi:10.1016/0003-9861(84)90507-1. PMID   6091559.
  133. Guranowski, Andrzej; Pawelkiewicz, Jerzy (1977). "Adenosylhomocysteinase from Yellow Lupin Seeds. Purification and Properties". European Journal of Biochemistry. 80 (2): 517–23. doi: 10.1111/j.1432-1033.1977.tb11907.x . PMID   923592.
  134. Kajander, EO; Raina, AM (1981). "Affinity-chromatographic purification of S-adenosyl-L-homocysteine hydrolase. Some properties of the enzyme from rat liver". The Biochemical Journal. 193 (2): 503–12. doi:10.1042/bj1930503. PMC   1162632 . PMID   7305945.
  135. 1 2 3 Saeki, Y; Ito, S; Shizuta, Y; Hayaishi, O; Kagamiyama, H; Wada, H (1977). "Subunit structure of biodegradative threonine deaminase". The Journal of Biological Chemistry. 252 (7): 2206–8. doi: 10.1016/S0021-9258(17)40542-4 . PMID   321452.
  136. 1 2 3 Phillips, A.T.; Wood, W.A. (1964). "Basis for AMP activation of "Biodegradative" threonine dehydrase from". Biochemical and Biophysical Research Communications. 15 (6): 530–535. doi:10.1016/0006-291X(64)90499-1.
  137. 1 2 3 Gerlt, JA; Rabinowitz, KW; Dunne, CP; Wood, WA (1973). "The mechanism of action of 5'-adenylic acid-activated threonine dehydrase. V. Relation between ligand-induced allosteric activation and the protomeroligomer interconversion". The Journal of Biological Chemistry. 248 (23): 8200–6. doi: 10.1016/S0021-9258(19)43214-6 . PMID   4584826.
  138. Addington, Adele K.; Johnson, David A. (1996). "Inactivation of Human Lung Tryptase: Evidence for a Re-Activatable Tetrameric Intermediate and Active Monomers". Biochemistry. 35 (42): 13511–8. doi:10.1021/bi960042t. PMID   8885830.
  139. Fajardo, Ignacio; Pejler, Gunnar (2003). "Formation of active monomers from tetrameric human β-tryptase". Biochemical Journal. 369 (3): 603–10. doi:10.1042/BJ20021418. PMC   1223112 . PMID   12387726.
  140. Fukuoka, Yoshihiro; Schwartz, Lawrence B. (2004). "Human β-Tryptase: Detection and Characterization of the Active Monomer and Prevention of Tetramer Reconstitution by Protease Inhibitors". Biochemistry. 43 (33): 10757–64. doi:10.1021/bi049486c. PMID   15311937.
  141. Fukuoka, Y; Schwartz, LB (2006). "The B12 anti-tryptase monoclonal antibody disrupts the tetrameric structure of heparin-stabilized beta-tryptase to form monomers that are inactive at neutral pH and active at acidic pH". Journal of Immunology. 176 (5): 3165–72. doi:10.4049/jimmunol.176.5.3165. PMC   1810230 . PMID   16493076.
  142. Fukuoka, Yoshihiro; Schwartz, Lawrence B. (2007). "Active monomers of human β-tryptase have expanded substrate specificities". International Immunopharmacology. 7 (14): 1900–8. doi:10.1016/j.intimp.2007.07.007. PMC   2278033 . PMID   18039527.
  143. Hallgren, J.; Spillmann, D; Pejler, G (2001). "Structural Requirements and Mechanism for Heparin-induced Activation of a Recombinant Mouse Mast Cell Tryptase, Mouse Mast Cell Protease-6. FORMATION OF ACTIVE TRYPTASE MONOMERS IN THE PRESENCE OF LOW MOLECULAR WEIGHT HEPARIN". Journal of Biological Chemistry. 276 (46): 42774–81. doi: 10.1074/jbc.M105531200 . PMID   11533057.
  144. Schechter, Norman M.; Choi, Eun-Jung; Selwood, Trevor; McCaslin, Darrell R. (2007). "Characterization of Three Distinct Catalytic Forms of Human Tryptase-β: Their Interrelationships and Relevance". Biochemistry. 46 (33): 9615–29. doi:10.1021/bi7004625. PMID   17655281.
  145. Schechter, Norman M.; Eng, Grace Y.; Selwood, Trevor; McCaslin, Darrell R. (1995). "Structural Changes Associated with the Spontaneous Inactivation of the Serine Proteinase Human Tryptase". Biochemistry. 34 (33): 10628–38. doi:10.1021/bi00033a038. PMID   7654717.
  146. Schwartz, Lawrence B. (1994). "[6] Tryptase: A mast cell serine protease". Proteolytic Enzymes: Serine and Cysteine Peptidases. Methods in Enzymology. Vol. 244. pp.  88–100. doi:10.1016/0076-6879(94)44008-5. ISBN   978-0-12-182145-6. PMID   7845247.
  147. Strik, Merel C. M.; Wolbink, Angela; Wouters, Dorine; Bladergroen, Bellinda A.; Verlaan, Angelique R.; van Houdt, Inge S.; Hijlkema, Sanne; Hack, C. Erik; et al. (2004). "Intracellular serpin SERPINB6 (PI6) is abundantly expressed by human mast cells and forms complexes with β-tryptase monomers". Blood. 103 (7): 2710–7. doi:10.1182/blood-2003-08-2981. PMID   14670919.
  148. 1 2 Kozik, Andrzej; Potempa, Jan; Travis, James (1998). "Spontaneous inactivation of human lung tryptase as probed by size-exclusion chromatography and chemical cross-linking: Dissociation of active tetrameric enzyme into inactive monomers is the primary event of the entire process". Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology. 1385 (1): 139–48. doi:10.1016/S0167-4838(98)00053-3. PMID   9630576.
  149. Alzani, R.; Cozzi, E.; Corti, A.; Temponi, M.; Trizio, D.; Gigli, M.; Rizzo, V. (1995). "Mechanism of suramin-induced deoligomerization of tumor necrosis factor .alpha". Biochemistry. 34 (19): 6344–50. doi:10.1021/bi00019a012. PMID   7756262.
  150. Corti, A; Fassina, G; Marcucci, F; Barbanti, E; Cassani, G (1992). "Oligomeric tumour necrosis factor alpha slowly converts into inactive forms at bioactive levels". The Biochemical Journal. 284 (3): 905–10. doi:10.1042/bj2840905. PMC   1132625 . PMID   1622406.
  151. Hlodan, Roman; Pain, Roger H. (1995). "The Folding and Assembly Pathway of Tumour Necrosis Factor TNFalpha, a Globular Trimeric Protein". European Journal of Biochemistry. 231 (2): 381–7. doi: 10.1111/j.1432-1033.1995.0381e.x . PMID   7635149.
  152. 1 2 3 4 Jensen, Kaj Frank; Mygind, Bente (1996). "Different Oligomeric States are Involved in the Allosteric Behavior of Uracil Phosphoribosyltransferase from Escherichia Coli". European Journal of Biochemistry. 240 (3): 637–45. doi:10.1111/j.1432-1033.1996.0637h.x. PMID   8856065.