Protein dynamics

Last updated

In molecular biology, proteins are generally thought to adopt unique structures determined by their amino acid sequences. However, proteins are not strictly static objects, but rather populate ensembles of (sometimes similar) conformations. Transitions between these states occur on a variety of length scales (tenths of angstroms to nm) and time scales (ns to s), and have been linked to functionally relevant phenomena such as allosteric signaling [1] and enzyme catalysis. [2]

Contents

The study of protein dynamics is most directly concerned with the transitions between these states, but can also involve the nature and equilibrium populations of the states themselves. These two perspectives—kinetics and thermodynamics, respectively—can be conceptually synthesized in an "energy landscape" paradigm: [3] highly populated states and the kinetics of transitions between them can be described by the depths of energy wells and the heights of energy barriers, respectively.

Kinesin walking on a microtubule. It is a molecular biological machine that uses protein domain dynamics on nanoscales Kinesin walking.gif
Kinesin walking on a microtubule. It is a molecular biological machine that uses protein domain dynamics on nanoscales

Local flexibility: atoms and residues

Portions of protein structures often deviate from the equilibrium state. Some such excursions are harmonic, such as stochastic fluctuations of chemical bonds and bond angles. Others are anharmonic, such as sidechains that jump between separate discrete energy minima, or rotamers. [4]

Evidence for local flexibility is often obtained from NMR spectroscopy. Flexible and potentially disordered regions of a protein can be detected using the random coil index. Flexibility in folded proteins can be identified by analyzing the spin relaxation of individual atoms in the protein. Flexibility can also be observed in very high-resolution electron density maps produced by X-ray crystallography, [5] particularly when diffraction data is collected at room temperature instead of the traditional cryogenic temperature (typically near 100 K). [6] Information on the frequency distribution and dynamics of local protein flexibility can be obtained using Raman and optical Kerr-effect spectroscopy [7] as well as anisotropic microspectroscopy [8] in the terahertz frequency domain.

Regional flexibility: intra-domain multi-residue coupling

A network of alternative conformations in catalase (Protein Data Bank code: 1gwe) with diverse properties. Multiple phenomena define the network: van der Waals interactions (blue dots and line segments) between sidechains, a hydrogen bond (dotted green line) through a partial-occupancy water (brown), coupling through the locally mobile backbone (black), and perhaps electrostatic forces between the Lys (green) and nearby polar residues (blue: Glu, yellow: Asp, purple: Ser). This particular network is distal from the active site and is therefore putatively not critical for function. Catalase diverse alternate conformation network.jpg
A network of alternative conformations in catalase (Protein Data Bank code: 1gwe) with diverse properties. Multiple phenomena define the network: van der Waals interactions (blue dots and line segments) between sidechains, a hydrogen bond (dotted green line) through a partial-occupancy water (brown), coupling through the locally mobile backbone (black), and perhaps electrostatic forces between the Lys (green) and nearby polar residues (blue: Glu, yellow: Asp, purple: Ser). This particular network is distal from the active site and is therefore putatively not critical for function.

Many residues are in close spatial proximity in protein structures. This is true for most residues that are contiguous in the primary sequence, but also for many that are distal in sequence yet are brought into contact in the final folded structure. Because of this proximity, these residue's energy landscapes become coupled based on various biophysical phenomena such as hydrogen bonds, ionic bonds, and van der Waals interactions (see figure).

Transitions between states for such sets of residues therefore become correlated. [9]

This is perhaps most obvious for surface-exposed loops, which often shift collectively to adopt different conformations in different crystal structures (see figure). However, coupled conformational heterogeneity is also sometimes evident in secondary structure. [10] For example, consecutive residues and residues offset by 4 in the primary sequence often interact in α helices. Also, residues offset by 2 in the primary sequence point their sidechains toward the same face of β sheets and are close enough to interact sterically, as are residues on adjacent strands of the same β sheet. Some of these conformational changes are induced by post-translational modifications in protein structure, such as phosphorylation and methylation. [10] [11]

An "ensemble" of 44 crystal structures of hen egg white lysozyme from the Protein Data Bank, showing that different crystallization conditions lead to different conformations for various surface-exposed loops and termini (red arrows). Hen egg white lysozyme PDB ensemble.jpg
An "ensemble" of 44 crystal structures of hen egg white lysozyme from the Protein Data Bank, showing that different crystallization conditions lead to different conformations for various surface-exposed loops and termini (red arrows).

When these coupled residues form pathways linking functionally important parts of a protein, they may participate in allosteric signaling. For example, when a molecule of oxygen binds to one subunit of the hemoglobin tetramer, that information is allosterically propagated to the other three subunits, thereby enhancing their affinity for oxygen. In this case, the coupled flexibility in hemoglobin allows for cooperative oxygen binding, which is physiologically useful because it allows rapid oxygen loading in lung tissue and rapid oxygen unloading in oxygen-deprived tissues (e.g. muscle).

Global flexibility: multiple domains

The presence of multiple domains in proteins gives rise to a great deal of flexibility and mobility, leading to protein domain dynamics. [1] Domain motions can be inferred by comparing different structures of a protein (as in Database of Molecular Motions), or they can be directly observed using spectra [12] [13] measured by neutron spin echo spectroscopy. They can also be suggested by sampling in extensive molecular dynamics trajectories [14] and principal component analysis. [15] Domain motions are important for:

One of the largest observed domain motions is the 'swivelling' mechanism in pyruvate phosphate dikinase. The phosphoinositide domain swivels between two states in order to bring a phosphate group from the active site of the nucleotide binding domain to that of the phosphoenolpyruvate/pyruvate domain. [23] The phosphate group is moved over a distance of 45 Å involving a domain motion of about 100 degrees around a single residue. In enzymes, the closure of one domain onto another captures a substrate by an induced fit, allowing the reaction to take place in a controlled way. A detailed analysis by Gerstein led to the classification of two basic types of domain motion; hinge and shear. [20] Only a relatively small portion of the chain, namely the inter-domain linker and side chains undergo significant conformational changes upon domain rearrangement. [24]

Hinge motions

Hinge motion in disordered activation domain in Trypsinogen (PDB ID: 2PTN). The hinges predicted using PACKMAN Hinge prediction are colored in blue (residues 23:28) and red (residues 175:182). The green colored region is the active site. Motion is generated using hdANM. 2PTN.gif
Hinge motion in disordered activation domain in Trypsinogen (PDB ID: 2PTN). The hinges predicted using PACKMAN Hinge prediction are colored in blue (residues 23:28) and red (residues 175:182). The green colored region is the active site. Motion is generated using hdANM.

A study by Hayward [25] found that the termini of α-helices and β-sheets form hinges in a large number of cases. Many hinges were found to involve two secondary structure elements acting like hinges of a door, allowing an opening and closing motion to occur. This can arise when two neighbouring strands within a β-sheet situated in one domain, diverge apart as they join the other domain. The two resulting termini then form the bending regions between the two domains. α-helices that preserve their hydrogen bonding network when bent are found to behave as mechanical hinges, storing `elastic energy' that drives the closure of domains for rapid capture of a substrate. [25] Khade et. al. worked on prediction of the hinges [26] in any conformation and further built an Elastic Network Model called hdANM [27] that can model those motions.

Helical to extended conformation

The interconversion of helical and extended conformations at the site of a domain boundary is not uncommon. In calmodulin, torsion angles change for five residues in the middle of a domain linking α-helix. The helix is split into two, almost perpendicular, smaller helices separated by four residues of an extended strand. [28] [29]

Shear motions

Shear motions involve a small sliding movement of domain interfaces, controlled by the amino acid side chains within the interface. Proteins displaying shear motions often have a layered architecture: stacking of secondary structures. The interdomain linker has merely the role of keeping the domains in close proximity.[ citation needed ]

Domain motion and functional dynamics in enzymes

The analysis of the internal dynamics of structurally different, but functionally similar enzymes has highlighted a common relationship between the positioning of the active site and the two principal protein sub-domains. In fact, for several members of the hydrolase superfamily, the catalytic site is located close to the interface separating the two principal quasi-rigid domains. [14] Such positioning appears instrumental for maintaining the precise geometry of the active site, while allowing for an appreciable functionally oriented modulation of the flanking regions resulting from the relative motion of the two sub-domains.[ citation needed ]

Implications for macromolecular evolution

Evidence suggests that protein dynamics are important for function, e.g. enzyme catalysis in dihydrofolate reductase (DHFR), yet they are also posited to facilitate the acquisition of new functions by molecular evolution. [30] This argument suggests that proteins have evolved to have stable, mostly unique folded structures, but the unavoidable residual flexibility leads to some degree of functional promiscuity, which can be amplified/harnessed/diverted by subsequent mutations.[ citation needed ] Research on promiscuous proteins within the BCL-2 family revealed that nanosecond-scale protein dynamics can play a crucial role in protein binding behaviour and thus promiscuity [31] .

However, there is growing awareness that intrinsically unstructured proteins are quite prevalent in eukaryotic genomes, [32] casting further doubt on the simplest interpretation of Anfinsen's dogma: "sequence determines structure (singular)". In effect, the new paradigm is characterized by the addition of two caveats: "sequence and cellular environment determine structural ensemble".

Related Research Articles

<span class="mw-page-title-main">Protein secondary structure</span> General three-dimensional form of local segments of proteins

Protein secondary structure is the local spatial conformation of the polypeptide backbone excluding the side chains. The two most common secondary structural elements are alpha helices and beta sheets, though beta turns and omega loops occur as well. Secondary structure elements typically spontaneously form as an intermediate before the protein folds into its three dimensional tertiary structure.

<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">Phenylalanine hydroxylase</span> Mammalian protein found in Homo sapiens

Phenylalanine hydroxylase (PAH) (EC 1.14.16.1) is an enzyme that catalyzes the hydroxylation of the aromatic side-chain of phenylalanine to generate tyrosine. PAH is one of three members of the biopterin-dependent aromatic amino acid hydroxylases, a class of monooxygenase that uses tetrahydrobiopterin (BH4, a pteridine cofactor) and a non-heme iron for catalysis. During the reaction, molecular oxygen is heterolytically cleaved with sequential incorporation of one oxygen atom into BH4 and phenylalanine substrate. In humans, mutations in its encoding gene, PAH, can lead to the metabolic disorder phenylketonuria.

<span class="mw-page-title-main">Protein structure prediction</span> Type of biological prediction

Protein structure prediction is the inference of the three-dimensional structure of a protein from its amino acid sequence—that is, the prediction of its secondary and tertiary structure from primary structure. Structure prediction is different from the inverse problem of protein design. Protein structure prediction is one of the most important goals pursued by computational biology; it is important in medicine and biotechnology.

<span class="mw-page-title-main">Lac repressor</span> DNA-binding protein

The lac repressor (LacI) is a DNA-binding protein that inhibits the expression of genes coding for proteins involved in the metabolism of lactose in bacteria. These genes are repressed when lactose is not available to the cell, ensuring that the bacterium only invests energy in the production of machinery necessary for uptake and utilization of lactose when lactose is present. When lactose becomes available, it is firstly converted into allolactose by β-Galactosidase (lacZ) in bacteria. The DNA binding ability of lac repressor bound with allolactose is inhibited due to allosteric regulation, thereby genes coding for proteins involved in lactose uptake and utilization can be expressed.

<span class="mw-page-title-main">Binding site</span> Molecule-specific coordinate bonding area in biological systems

In biochemistry and molecular biology, a binding site is a region on a macromolecule such as a protein that binds to another molecule with specificity. The binding partner of the macromolecule is often referred to as a ligand. Ligands may include other proteins, enzyme substrates, second messengers, hormones, or allosteric modulators. The binding event is often, but not always, accompanied by a conformational change that alters the protein's function. Binding to protein binding sites is most often reversible, but can also be covalent reversible or irreversible.

<span class="mw-page-title-main">Protein structure</span> Three-dimensional arrangement of atoms in an amino acid-chain molecule

Protein structure is the three-dimensional arrangement of atoms in an amino acid-chain molecule. Proteins are polymers – specifically polypeptides – formed from sequences of amino acids, which are the monomers of the polymer. A single amino acid monomer may also be called a residue, which indicates a repeating unit of a polymer. Proteins form by amino acids undergoing condensation reactions, in which the amino acids lose one water molecule per reaction in order to attach to one another with a peptide bond. By convention, a chain under 30 amino acids is often identified as a peptide, rather than a protein. To be able to perform their biological function, proteins fold into one or more specific spatial conformations driven by a number of non-covalent interactions, such as hydrogen bonding, ionic interactions, Van der Waals forces, and hydrophobic packing. To understand the functions of proteins at a molecular level, it is often necessary to determine their three-dimensional structure. This is the topic of the scientific field of structural biology, which employs techniques such as X-ray crystallography, NMR spectroscopy, cryo-electron microscopy (cryo-EM) and dual polarisation interferometry, to determine the structure of proteins.

<span class="mw-page-title-main">Intrinsically disordered proteins</span> Protein without a fixed 3D structure

In molecular biology, an intrinsically disordered protein (IDP) is a protein that lacks a fixed or ordered three-dimensional structure, typically in the absence of its macromolecular interaction partners, such as other proteins or RNA. IDPs range from fully unstructured to partially structured and include random coil, molten globule-like aggregates, or flexible linkers in large multi-domain proteins. They are sometimes considered as a separate class of proteins along with globular, fibrous and membrane proteins.

<span class="mw-page-title-main">Conformational change</span> Change in the shape of a macromolecule, often induced by environmental factors

In biochemistry, a conformational change is a change in the shape of a macromolecule, often induced by environmental factors.

A turn is an element of secondary structure in proteins where the polypeptide chain reverses its overall direction.

<span class="mw-page-title-main">Phosphoglycerate kinase</span> Enzyme

Phosphoglycerate kinase is an enzyme that catalyzes the reversible transfer of a phosphate group from 1,3-bisphosphoglycerate (1,3-BPG) to ADP producing 3-phosphoglycerate (3-PG) and ATP :

Allosteric enzymes are enzymes that change their conformational ensemble upon binding of an effector which results in an apparent change in binding affinity at a different ligand binding site. This "action at a distance" through binding of one ligand affecting the binding of another at a distinctly different site, is the essence of the allosteric concept. Allostery plays a crucial role in many fundamental biological processes, including but not limited to cell signaling and the regulation of metabolism. Allosteric enzymes need not be oligomers as previously thought, and in fact many systems have demonstrated allostery within single enzymes. In biochemistry, allosteric regulation is the regulation of a protein by binding an effector molecule at a site other than the enzyme's active site.

<span class="mw-page-title-main">Molecular biophysics</span> Interdisciplinary research area

Molecular biophysics is a rapidly evolving interdisciplinary area of research that combines concepts in physics, chemistry, engineering, mathematics and biology. It seeks to understand biomolecular systems and explain biological function in terms of molecular structure, structural organization, and dynamic behaviour at various levels of complexity. This discipline covers topics such as the measurement of molecular forces, molecular associations, allosteric interactions, Brownian motion, and cable theory. Additional areas of study can be found on Outline of Biophysics. The discipline has required development of specialized equipment and procedures capable of imaging and manipulating minute living structures, as well as novel experimental approaches.

<span class="mw-page-title-main">Protein domain</span> Self-stable region of a proteins chain that folds independently from the rest

In molecular biology, a protein domain is a region of a protein's polypeptide chain that is self-stabilizing and that folds independently from the rest. Each domain forms a compact folded three-dimensional structure. Many proteins consist of several domains, and a domain may appear in a variety of different proteins. Molecular evolution uses domains as building blocks and these may be recombined in different arrangements to create proteins with different functions. In general, domains vary in length from between about 50 amino acids up to 250 amino acids in length. The shortest domains, such as zinc fingers, are stabilized by metal ions or disulfide bridges. Domains often form functional units, such as the calcium-binding EF hand domain of calmodulin. Because they are independently stable, domains can be "swapped" by genetic engineering between one protein and another to make chimeric proteins.

<span class="mw-page-title-main">Gaussian network model</span>

The Gaussian network model (GNM) is a representation of a biological macromolecule as an elastic mass-and-spring network to study, understand, and characterize the mechanical aspects of its long-time large-scale dynamics. The model has a wide range of applications from small proteins such as enzymes composed of a single domain, to large macromolecular assemblies such as a ribosome or a viral capsid. Protein domain dynamics plays key roles in a multitude of molecular recognition and cell signalling processes. Protein domains, connected by intrinsically disordered flexible linker domains, induce long-range allostery via protein domain dynamics. The resultant dynamic modes cannot be generally predicted from static structures of either the entire protein or individual domains.

<span class="mw-page-title-main">Fuzzy complex</span>

Fuzzy complexes are protein complexes, where structural ambiguity or multiplicity exists and is required for biological function. Alteration, truncation or removal of conformationally ambiguous regions impacts the activity of the corresponding complex. Fuzzy complexes are generally formed by intrinsically disordered proteins. Structural multiplicity usually underlies functional multiplicity of protein complexes following a fuzzy logic. Distinct binding modes of the nucleosome are also regarded as a special case of fuzziness.

<span class="mw-page-title-main">KcsA potassium channel</span> Prokaryotic potassium ion channel

KcsA (Kchannel of streptomyces A) is a prokaryotic potassium channel from the soil bacterium Streptomyces lividans that has been studied extensively in ion channel research. The pH activated protein possesses two transmembrane segments and a highly selective pore region, responsible for the gating and shuttling of K+ ions out of the cell. The amino acid sequence found in the selectivity filter of KcsA is highly conserved among both prokaryotic and eukaryotic K+ voltage channels; as a result, research on KcsA has provided important structural and mechanistic insight on the molecular basis for K+ ion selection and conduction. As one of the most studied ion channels to this day, KcsA is a template for research on K+ channel function and its elucidated structure underlies computational modeling of channel dynamics for both prokaryotic and eukaryotic species.

A protein superfamily is the largest grouping (clade) of proteins for which common ancestry can be inferred. Usually this common ancestry is inferred from structural alignment and mechanistic similarity, even if no sequence similarity is evident. Sequence homology can then be deduced even if not apparent. Superfamilies typically contain several protein families which show sequence similarity within each family. The term protein clan is commonly used for protease and glycosyl hydrolases superfamilies based on the MEROPS and CAZy classification systems.

<span class="mw-page-title-main">Niche (protein structural motif)</span>

In the area of protein structural motifs, niches are three or four amino acid residue features in which main-chain CO groups are bridged by positively charged or δ+ groups. The δ+ groups include groups with two hydrogen bond donor atoms such as NH2 groups and water molecules. In typical proteins, 7% of amino acid residues belong to niches bound to a δ+ group, while another 7% have the conformation but no single cationic bridging group is detected.

References

  1. 1 2 Bu Z, Callaway DJ (2011). "Proteins move! Protein dynamics and long-range allostery in cell signaling". In Donev R (ed.). Protein Structure and Diseases. Advances in Protein Chemistry and Structural Biology. Vol. 83. Academic Press. pp. 163–221. doi:10.1016/B978-0-12-381262-9.00005-7. ISBN   9780123812629. PMID   21570668.
  2. Fraser JS, Clarkson MW, Degnan SC, Erion R, Kern D, Alber T (Dec 2009). "Hidden alternative structures of proline isomerase essential for catalysis". Nature. 462 (7273): 669–673. Bibcode:2009Natur.462..669F. doi:10.1038/nature08615. PMC   2805857 . PMID   19956261.
  3. Frauenfelder H, Sligar SG, Wolynes PG (Dec 1991). "The energy landscapes and motions of proteins". Science. 254 (5038): 1598–1603. Bibcode:1991Sci...254.1598F. doi:10.1126/science.1749933. PMID   1749933.
  4. Dunbrack, Roland L (August 2002). "Rotamer Libraries in the 21st Century". Current Opinion in Structural Biology. 12 (4): 431–440. doi:10.1016/s0959-440x(02)00344-5. PMID   12163064.
  5. Davis IW, Arendall WB, Richardson DC, Richardson JS (Feb 2006). "The backrub motion: how protein backbone shrugs when a sidechain dances". Structure. 14 (2): 265–274. doi: 10.1016/j.str.2005.10.007 . PMID   16472746.
  6. Fraser JS, van den Bedem H, Samelson AJ, Lang PT, Holton JM, Echols N, Alber T (Sep 2011). "Accessing protein conformational ensembles using room-temperature X-ray crystallography". Proceedings of the National Academy of Sciences of the United States of America. 108 (39): 16247–16252. Bibcode:2011PNAS..10816247F. doi: 10.1073/pnas.1111325108 . PMC   3182744 . PMID   21918110.
  7. Turton DA, Senn HM, Harwood T, Lapthorn AJ, Ellis EM, Wynne K (June 2014). "Terahertz underdamped vibrational motion governs protein-ligand binding in solution". Nature Communications. 5: 3999. Bibcode:2014NatCo...5.3999T. doi: 10.1038/ncomms4999 . PMID   24893252.
  8. Acbas, G.; Niessen, K. A.; Snell, E. H.; Markelz, A. G. (2014). "Protein Optical measurements of long-range protein vibrations". Nature Communications. 5: 3076. doi: 10.1038/ncomms4076 . PMID   24430203.
  9. Bu Z, Cook J, Callaway DJ (Sep 2001). "Dynamic regimes and correlated structural dynamics in native and denatured alpha-lactalbumin". Journal of Molecular Biology. 312 (4): 865–873. doi:10.1006/jmbi.2001.5006. PMID   11575938.
  10. 1 2 Costa CH, Oliveira AR, Dos Santos AM, da Costa KS, Lima AH, Alves CN, Lameira J (October 2019). "Computational study of conformational changes in human 3-hydroxy-3-methylglutaryl coenzyme reductase induced by substrate binding". Journal of Biomolecular Structure & Dynamics. 37 (16): 4374–4383. doi:10.1080/07391102.2018.1549508. PMID   30470158. S2CID   53717806.
  11. Groban ES, Narayanan A, Jacobson MP (April 2006). Shakhnovich E (ed.). "Conformational changes in protein loops and helices induced by post-translational phosphorylation". PLOS Computational Biology. 2 (4): e32. Bibcode:2006PLSCB...2...32G. doi: 10.1371/journal.pcbi.0020032 . PMC   1440919 . PMID   16628247.
  12. Farago B, Li J, Cornilescu G, Callaway DJ, Bu Z (Nov 2010). "Activation of nanoscale allosteric protein domain motion revealed by neutron spin echo spectroscopy". Biophysical Journal. 99 (10): 3473–3482. Bibcode:2010BpJ....99.3473F. doi:10.1016/j.bpj.2010.09.058. PMC   2980739 . PMID   21081097.
  13. Bu Z, Biehl R, Monkenbusch M, Richter D, Callaway DJ (Dec 2005). "Coupled protein domain motion in Taq polymerase revealed by neutron spin-echo spectroscopy" (PDF). Proceedings of the National Academy of Sciences of the United States of America. 102 (49): 17646–17651. Bibcode:2005PNAS..10217646B. doi: 10.1073/pnas.0503388102 . PMC   1345721 . PMID   16306270.
  14. 1 2 Potestio R, Pontiggia F, Micheletti C (Jun 2009). "Coarse-grained description of protein internal dynamics: an optimal strategy for decomposing proteins in rigid subunits". Biophysical Journal. 96 (12): 4993–5002. Bibcode:2009BpJ....96.4993P. doi:10.1016/j.bpj.2009.03.051. PMC   2712024 . PMID   19527659.
  15. Baron R, Vellore NA (Jul 2012). "LSD1/CoREST is an allosteric nanoscale clamp regulated by H3-histone-tail molecular recognition". Proceedings of the National Academy of Sciences of the United States of America. 109 (31): 12509–14. Bibcode:2012PNAS..10912509B. doi: 10.1073/pnas.1207892109 . PMC   3411975 . PMID   22802671.
  16. Ponte-Sucre A, ed. (2009). ABC Transporters in Microorganisms. Caister Academic. ISBN   978-1-904455-49-3.
  17. Kamerlin SC, Warshel A (May 2010). "At the dawn of the 21st century: Is dynamics the missing link for understanding enzyme catalysis?". Proteins. 78 (6): 1339–75. doi:10.1002/prot.22654. PMC   2841229 . PMID   20099310.
  18. Howard J (2001). Mechanics of motor proteins and the cytoskeleton (1st ed.). Sunderland,MA: Sinauer Associates. ISBN   9780878933334.
  19. Callaway DJ, Matsui T, Weiss T, Stingaciu LR, Stanley CB, Heller WT, Bu Z (Apr 2017). "Controllable Activation of Nanoscale Dynamics in a Disordered Protein Alters Binding Kinetics". Journal of Molecular Biology. 429 (7): 987–998. doi:10.1016/j.jmb.2017.03.003. PMC   5399307 . PMID   28285124.
  20. 1 2 Gerstein M, Lesk AM, Chothia C (Jun 1994). "Structural mechanisms for domain movements in proteins". Biochemistry. 33 (22): 6739–49. doi:10.1021/bi00188a001. PMID   8204609.
  21. Nicholl ID, Matsui T, Weiss TM, Stanley CB, Heller WT, Martel A, Farago B, Callaway DJ, Bu Z (Aug 21, 2018). "Alpha-catenin structure and nanoscale dynamics in solution and in complex with F-actin". Biophysical Journal. 115 (4): 642–654. Bibcode:2018BpJ...115..642N. doi:10.1016/j.bpj.2018.07.005. hdl:2436/621755. PMC   6104293 . PMID   30037495.
  22. Voet D (2011). Biochemistry. Voet, Judith G. (4th ed.). Hoboken, NJ: John Wiley & Sons. ISBN   9780470570951. OCLC   690489261.
  23. Herzberg O, Chen CC, Kapadia G, McGuire M, Carroll LJ, Noh SJ, Dunaway-Mariano D (Apr 1996). "Swiveling-domain mechanism for enzymatic phosphotransfer between remote reaction sites". Proceedings of the National Academy of Sciences of the United States of America. 93 (7): 2652–7. Bibcode:1996PNAS...93.2652H. doi: 10.1073/pnas.93.7.2652 . PMC   39685 . PMID   8610096.
  24. Janin J, Wodak SJ (1983). "Structural domains in proteins and their role in the dynamics of protein function". Progress in Biophysics and Molecular Biology. 42 (1): 21–78. doi: 10.1016/0079-6107(83)90003-2 . PMID   6353481.
  25. 1 2 Hayward S (Sep 1999). "Structural principles governing domain motions in proteins". Proteins. 36 (4): 425–35. doi:10.1002/(SICI)1097-0134(19990901)36:4<425::AID-PROT6>3.0.CO;2-S. PMID   10450084. S2CID   29808315.
  26. Khade, Pranav M.; Kumar, Ambuj; Jernigan, Robert L. (2020-01-17). "Characterizing and Predicting Protein Hinges for Mechanistic Insight". Journal of Molecular Biology. 432 (2): 508–522. doi:10.1016/j.jmb.2019.11.018. ISSN   1089-8638. PMC   7029793 . PMID   31786268.
  27. Khade, Pranav M.; Scaramozzino, Domenico; Kumar, Ambuj; Lacidogna, Giuseppe; Carpinteri, Alberto; Jernigan, Robert L. (2021-11-16). "hdANM: a new comprehensive dynamics model for protein hinges". Biophysical Journal. 120 (22): 4955–4965. doi:10.1016/j.bpj.2021.10.017. ISSN   1542-0086. PMC   8633836 . PMID   34687719.
  28. Meador WE, Means AR, Quiocho FA (Aug 1992). "Target enzyme recognition by calmodulin: 2.4 A structure of a calmodulin-peptide complex". Science. 257 (5074): 1251–1255. Bibcode:1992Sci...257.1251M. doi:10.1126/science.1519061. PMID   1519061.
  29. Ikura M, Clore GM, Gronenborn AM, Zhu G, Klee CB, Bax A (May 1992). "Solution structure of a calmodulin-target peptide complex by multidimensional NMR". Science. 256 (5057): 632–638. Bibcode:1992Sci...256..632I. doi:10.1126/science.1585175. PMID   1585175.
  30. Tokuriki N, Tawfik DS (Apr 2009). "Protein dynamism and evolvability". Science. 324 (5924): 203–207. Bibcode:2009Sci...324..203T. doi:10.1126/science.1169375. PMID   19359577. S2CID   28576496.
  31. Heckmeier, Philipp J.; Ruf, Jeannette; Janković, Brankica G.; Hamm, Peter (7 March 2023). "MCL-1 promiscuity and the structural resilience of its binding partners". The Journal of Chemical Physics. 158 (9). doi:10.1063/5.0137239.
  32. Dyson HJ, Wright PE (Mar 2005). "Intrinsically unstructured proteins and their functions". Nature Reviews Molecular Cell Biology. 6 (3): 197–208. doi:10.1038/nrm1589. PMID   15738986. S2CID   18068406.