Protein quinary structure

Last updated

Protein quinary structure refers to the features of protein surfaces that are shaped by evolutionary adaptation to the physiological context of living cells. [1] [2] [3] [4] Quinary structure is thus the fifth level of protein complexity, additional to protein primary, secondary, tertiary and quaternary structures. As opposed to the first four levels of protein structure, which are relevant to isolated proteins in dilute conditions, quinary structure emerges from the crowdedness of the cellular context, [5] in which transient encounters among macromolecules are constantly occurring.

Contents

In order to perform their functions, proteins often need to find a specific counterpart to which they will bind in a relatively long encounter. In a very crowded cytosol, in which proteins engage in a vast and complex network of attracting and repelling interactions, such search becomes challenging, because it involves sampling a huge space of possible partners, of which very few will be productive. A solution to this challenge requires that proteins spend as little time as possible on each encounter, so that they can explore a larger number of surfaces, while simultaneously making this interaction as intimate as possible, so if they do come across the right partner, they will not miss it. [6] In this sense, quinary structure is the result of a series of adaptations present in protein surfaces, which allow proteins to navigate the complexity of the cellular environment.

Early observations

With the sense with which it is used today, the term quinary structure first appeared in the work of McConkey, in 1989. [7] In his work, McConkey runs 2D electrophoresis gels on the total protein content of hamster (CHO) and human (HeLa) cells. In a 2D electrophoresis gel experiment, the coordinates of a protein depend on its molecular weight and its isoelectric point. Given the evolutionary distance between humans and hamsters, and considering evolutionary rates typical of mammals, one would expect a large number of substitutions to have occurred between hamsters and humans, a fraction of which would involve acidic (aspartate and glutamate) and basic (arginine and lysine) residues, resulting in changes in the isoelectric point of many proteins. Strikingly, hamster and human cells yielded almost identical fingerprints in the experiment, implying that many fewer of those substitutions actually took place. McConkey suggested in that paper [7] that the reason why the proteins of humans and hamsters had not diverged as much he anticipated was that an additional selective pressure must have been related to the many non-specific “interactions that are inherently transient” experienced by proteins in the cytoplasm and which “constitute the fifth level of protein organization”.

Protein interactions and quinary structure

Despite the crudeness of McConkey's experiment, his interpretation of the results have proved to be accurate. Rather than simply being hydrophilic, protein surfaces must have carefully been modulated by evolution and adapted to this network of weak interactions, often called quinary interactions. It is important to note that protein-protein interactions responsible for the emergence of quinary structure are fundamentally different from specific protein encounters. The latter are the result of relatively high-stability binding, often linked to functionally meaningful events –many of which have already been described [8] – while the former are often interpreted as some background noise of physiologically unproductive misinteractions that complicate the interpretation of protein networks and need to be avoided, so that normal cellular functions can proceed. [9] [10] [11]

The transient nature of these protein encounters complicates the study of quinary structure. Indeed, the interactions responsible for this upper level of protein organisation are weak and short-lived, and hence would not produce protein-protein complexes that could be isolated by conventional biochemical methods. Therefore, quinary structure can only be understood in vivo. [12]

In-cell NMR and quinary structure

In-cell NMR is an experimental technique prominent in the research field of protein quinary structure. The physical principle of in-cell NMR measurements is identical to that of conventional protein NMR, but the experiments rely on expressing high concentrations of the probe protein, which should remain soluble and contained in the cellular space; which introduces additional difficulties and limitations. However, these experiments provide critical insights about the cross-talk between a probe protein and the intracellular environment.

Early attempts at using in-cell NMR to study protein quinary structure were hindered by a limitation caused by the very phenomenon they were trying to understand. Many probe proteins tested in these experiments turned out to produce broad signals, near the detection limit of the method, when measured inside cells of Escherichia coli. In particular, these proteins seemed to tumble as if they had molecular weights much larger than those corresponding to their size. These observations seemed to indicate that the proteins were sticking to other macromolecules, which would have led to poor relaxation properties [13]

Other in-cell NMR experiments showed that single amino acid changes of surface residues could be used to consistently modulate the tumbling of three different proteins inside bacterial cells. [14] Charged and hydrophobic residues were shown to have the largest impact in protein intracellular mobility. In particular, more negatively charged proteins would tumble faster in comparison with near-null or positively charged proteins. In contrast, the presence of many hydrophobic residues in the protein surface would slow down protein intracellular tumbling. Protein dipole moment, a measure of charge separation across the protein, was shown to have a significant contribution to protein mobility, where high dipole moments would correlate with slower tumbling.

Related Research Articles

<span class="mw-page-title-main">Chemotaxis</span> Movement of an organism or entity in response to a chemical stimulus

Chemotaxis is the movement of an organism or entity in response to a chemical stimulus. Somatic cells, bacteria, and other single-cell or multicellular organisms direct their movements according to certain chemicals in their environment. This is important for bacteria to find food by swimming toward the highest concentration of food molecules, or to flee from poisons. In multicellular organisms, chemotaxis is critical to early development and development as well as in normal function and health. In addition, it has been recognized that mechanisms that allow chemotaxis in animals can be subverted during cancer metastasis, and the aberrant change of the overall property of these networks, which control chemotaxis, can lead to carcinogenesis. The aberrant chemotaxis of leukocytes and lymphocytes also contribute to inflammatory diseases such as atherosclerosis, asthma, and arthritis. Sub-cellular components, such as the polarity patch generated by mating yeast, may also display chemotactic behavior.

<span class="mw-page-title-main">Protein</span> Biomolecule consisting of chains of amino acid residues

Proteins are large biomolecules and macromolecules that comprise one or more long chains of amino acid residues. Proteins perform a vast array of functions within organisms, including catalysing metabolic reactions, DNA replication, responding to stimuli, providing structure to cells and organisms, and transporting molecules from one location to another. Proteins differ from one another primarily in their sequence of amino acids, which is dictated by the nucleotide sequence of their genes, and which usually results in protein folding into a specific 3D structure that determines its activity.

<span class="mw-page-title-main">Protein folding</span> Change of a linear protein chain to a 3D structure

Protein folding is the physical process by which a protein, after synthesis by a ribosome as a linear chain of amino acids, changes from an unstable random coil into a more ordered three-dimensional structure. This structure permits the protein to become biologically functional.

<span class="mw-page-title-main">Peripheral membrane protein</span> Membrane proteins that adhere temporarily to membranes with which they are associated

Peripheral membrane proteins, or extrinsic membrane proteins, are membrane proteins that adhere only temporarily to the biological membrane with which they are associated. These proteins attach to integral membrane proteins, or penetrate the peripheral regions of the lipid bilayer. The regulatory protein subunits of many ion channels and transmembrane receptors, for example, may be defined as peripheral membrane proteins. In contrast to integral membrane proteins, peripheral membrane proteins tend to collect in the water-soluble component, or fraction, of all the proteins extracted during a protein purification procedure. Proteins with GPI anchors are an exception to this rule and can have purification properties similar to those of integral membrane proteins.

<span class="mw-page-title-main">Aspartate carbamoyltransferase</span> Protein family

Aspartate carbamoyltransferase catalyzes the first step in the pyrimidine biosynthetic pathway.

<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–protein interaction</span> Physical interactions and constructions between multiple proteins

Protein–protein interactions (PPIs) are physical contacts of high specificity established between two or more protein molecules as a result of biochemical events steered by interactions that include electrostatic forces, hydrogen bonding and the hydrophobic effect. Many are physical contacts with molecular associations between chains that occur in a cell or in a living organism in a specific biomolecular context.

<span class="mw-page-title-main">M2 proton channel</span>

The Matrix-2 (M2) protein is a proton-selective viroporin, integral in the viral envelope of the influenza A virus. The channel itself is a homotetramer, where the units are helices stabilized by two disulfide bonds, and is activated by low pH. The M2 protein is encoded on the seventh RNA segment together with the M1 protein. Proton conductance by the M2 protein in influenza A is essential for viral replication.

<span class="mw-page-title-main">Residual dipolar coupling</span>

The residual dipolar coupling between two spins in a molecule occurs if the molecules in solution exhibit a partial alignment leading to an incomplete averaging of spatially anisotropic dipolar couplings.

<span class="mw-page-title-main">Gla domain</span>

Vitamin K-dependent carboxylation/gamma-carboxyglutamic (GLA) domain is a protein domain that contains post-translational modifications of many glutamate residues by vitamin K-dependent carboxylation to form γ-carboxyglutamate (Gla). Proteins with this domain are known informally as Gla proteins. The Gla residues are responsible for the high-affinity binding of calcium ions.

<span class="mw-page-title-main">Nucleoporin 214</span> Protein-coding gene in the species Homo sapiens

Nucleoporin 214 (Nup2014) is a protein that in humans is encoded by the NUP214 gene.

<span class="mw-page-title-main">CLIC1</span> Protein-coding gene in the species Homo sapiens

Chloride intracellular channel protein 1 is a protein that in humans is encoded by the CLIC1 gene.

<span class="mw-page-title-main">Heat-stable enterotoxin</span> Class of bacterial toxins

Heat-stable enterotoxins (STs) are secretory peptides produced by some bacterial strains, such as enterotoxigenic Escherichia coli which are in general toxic to animals.

<span class="mw-page-title-main">CBX5 (gene)</span> Protein-coding gene in humans

Chromobox protein homolog 5 is a protein that in humans is encoded by the CBX5 gene. It is a highly conserved, non-histone protein part of the heterochromatin family. The protein itself is more commonly called HP1α. Heterochromatin protein-1 (HP1) has an N-terminal domain that acts on methylated lysines residues leading to epigenetic repression. The C-terminal of this protein has a chromo shadow-domain (CSD) that is responsible for homodimerizing, as well as interacting with a variety of chromatin-associated, non-histone proteins.

<span class="mw-page-title-main">Macromolecular crowding</span>

The phenomenon of macromolecular crowding alters the properties of molecules in a solution when high concentrations of macromolecules such as proteins are present. Such conditions occur routinely in living cells; for instance, the cytosol of Escherichia coli contains about 300–400 mg/ml of macromolecules. Crowding occurs since these high concentrations of macromolecules reduce the volume of solvent available for other molecules in the solution, which has the result of increasing their effective concentrations. Crowding can promote formation of a biomolecular condensate by colloidal phase separation.

Tat (HIV)

In molecular biology, Tat is a protein that is encoded for by the tat gene in HIV-1. Tat is a regulatory protein that drastically enhances the efficiency of viral transcription. Tat stands for "Trans-Activator of Transcription". The protein consists of between 86 and 101 amino acids depending on the subtype. Tat vastly increases the level of transcription of the HIV dsDNA. Before Tat is present, a small number of RNA transcripts will be made, which allow the Tat protein to be produced. Tat then binds to cellular factors and mediates their phosphorylation, resulting in increased transcription of all HIV genes, providing a positive feedback cycle. This in turn allows HIV to have an explosive response once a threshold amount of Tat is produced, a useful tool for defeating the body's response.

Experimental approaches of determining the structure of nucleic acids, such as RNA and DNA, can be largely classified into biophysical and biochemical methods. Biophysical methods use the fundamental physical properties of molecules for structure determination, including X-ray crystallography, NMR and cryo-EM. Biochemical methods exploit the chemical properties of nucleic acids using specific reagents and conditions to assay the structure of nucleic acids. Such methods may involve chemical probing with specific reagents, or rely on native or analogue chemistry. Different experimental approaches have unique merits and are suitable for different experimental purposes.

<span class="mw-page-title-main">Ammonia transporter</span>

Ammonia transporters are structurally related membrane transport proteins called Amt proteins in bacteria and plants, methylammonium/ammonium permeases (MEPs) in yeast, or Rhesus (Rh) proteins in chordates. In humans, the RhAG, RhBG, and RhCG Rhesus proteins constitute solute carrier family 42 whilst RhD and RhCE form the Rh blood group system. The three-dimensional structure of the ammonia transport protein AmtB from Escherichia coli has been determined by x-ray crystallography revealing a hydrophobic ammonia channel. The human RhCG ammonia transporter was found to have a similar ammonia-conducting channel structure. It was proposed that the erythrocyte Rh complex is a heterotrimer of RhAG, RhD, and RhCE subunits in which RhD and RhCE might play roles in anchoring the ammonia-conducting RhAG subunit to the cytoskeleton. Based on reconstitution experiments, purified RhCG subunits alone can function to transport ammonia. RhCG is required for normal acid excretion by the mouse kidney and epididymis.

<span class="mw-page-title-main">G. Marius Clore</span> Molecular biophysicist, structural biologist

G. Marius Clore MAE, FRSC, FMedSci, FRS is a British-born, Anglo-American molecular biophysicist and structural biologist. He was born in London, U.K. and is a dual U.S./U.K. Citizen. He is a Member of the National Academy of Sciences, a Fellow of the Royal Society, a Fellow of the Academy of Medical Sciences, a Fellow of the American Academy of Arts and Sciences, a NIH Distinguished Investigator, and the Chief of the Molecular and Structural Biophysics Section in the Laboratory of Chemical Physics of the National Institute of Diabetes and Digestive and Kidney Diseases at the U.S. National Institutes of Health. He is known for his foundational work in three-dimensional protein and nucleic acid structure determination by biomolecular NMR spectroscopy, for advancing experimental approaches to the study of large macromolecules and their complexes by NMR, and for developing NMR-based methods to study rare conformational states in protein-nucleic acid and protein-protein recognition. Clore's discovery of previously undetectable, functionally significant, rare transient states of macromolecules has yielded fundamental new insights into the mechanisms of important biological processes, and in particular the significance of weak interactions and the mechanisms whereby the opposing constraints of speed and specificity are optimized. Further, Clore's work opens up a new era of pharmacology and drug design as it is now possible to target structures and conformations that have been heretofore unseen.

<span class="mw-page-title-main">Mei Hong (chemist)</span> Chinese-American chemist

Mei Hong is a Chinese-American biophysical chemist and professor of chemistry at the Massachusetts Institute of Technology. She is known for her creative development and application of solid-state nuclear magnetic resonance (ssNMR) spectroscopy to elucidate the structures and mechanisms of membrane proteins, plant cell walls, and amyloid proteins. She has received a number of recognitions for her work, including the American Chemical Society Nakanishi Prize in 2021, Günther Laukien Prize in 2014, the Protein Society Young Investigator award in 2012, and the American Chemical Society’s Pure Chemistry award in 2003.

References

  1. Cohen, Rachel D.; Pielak, Gary J. (2016). "Electrostatic Contributions to Protein Quinary Structure". Journal of the American Chemical Society . 138 (40): 13139–13142. doi:10.1021/jacs.6b07323. PMID   27676610.
  2. Edelstein, S. J. (October 1980). "Patterns in the quinary structures of proteins. Plasticity and inequivalence of individual molecules in helical arrays of sickle cell hemoglobin and tubulin". Biophysical Journal. 32 (1): 347–360. Bibcode:1980BpJ....32..347E. doi:10.1016/S0006-3495(80)84961-7. PMC   1327314 . PMID   7248453.
  3. "Probing Protein Quinary Interactions by in-cell NMR". ResearchGate. Retrieved 2019-09-02.
  4. Shekhtman, Alexander; Burz, David S.; DeMott, Christopher; Breindel, Leonard (2018). "Real-Time In-Cell Nuclear Magnetic Resonance: Ribosome-Targeted Antibiotics Modulate Quinary Protein Interactions". Biochemistry. 57 (5). U.S.: United States Department of Agriculture: 540–546. doi:10.1021/acs.biochem.7b00938. PMC   5801172 . PMID   29266932 . Retrieved 2019-09-02.
  5. Danielsson, J.; Oliveberg, M. (2017). "Comparing protein behaviour in vitro and in vivo, what does the data really tell us?". Current Opinion in Structural Biology. 42: 129–135. doi:10.1016/j.sbi.2017.01.002. PMID   28126529.
  6. Jacek T. Mika; Bert Poolman (2011). "Macromolecule diffusion and confinement in prokaryotic cells". Current Opinion in Biotechnology. 22 (1): 117–126. doi:10.1016/j.copbio.2010.09.009. PMID   20952181.
  7. 1 2 McConkey, E. H. (1989). "Molecular evolution, intracellular organization, and the quinary structure of proteins". Proceedings of the National Academy of Sciences of the United States of America. 79 (10): 3236–3240. doi: 10.1073/pnas.79.10.3236 . PMC   346390 . PMID   6954476.
  8. Wlodarski, T.; Zagrovic, B. (2009). "Conformational selection and induced fit mechanism underlie specificity in noncovalent interactions with ubiquitin". Proceedings of the National Academy of Sciences of the United States of America. 106 (46): 3236–3240. Bibcode:2009PNAS..10619346W. doi: 10.1073/pnas.0906966106 . PMC   2780739 . PMID   19887638.
  9. Schreiber, G.; Fersht, A. R. (1996). "Rapid, electrostatically assisted association of proteins". Nature Structural Biology. 3 (5): 427–431. doi:10.1038/nsb0596-427. PMID   8612072. S2CID   25318867.
  10. Deeds, E. J.; Ashenberg, O.; Shakhnovich, E. I. (2006). "From The Cover: A simple physical model for scaling in protein-protein interaction networks". Proceedings of the National Academy of Sciences of the United States of America. 103 (2): 311–316. arXiv: q-bio/0509001 . Bibcode:2006PNAS..103..311D. doi: 10.1073/pnas.0509715102 . PMC   1326177 . PMID   16384916.
  11. Jian-Rong Yang; Ben-Yang Liao; Shi-Mei Zhuang; Jianzhi Zhang (2012). "Protein misinteraction avoidance causes highly expressed proteins to evolve slowly". Proceedings of the National Academy of Sciences of the United States of America. 109 (14): E831–E840. doi: 10.1073/pnas.1117408109 . PMC   3325723 . PMID   22416125.
  12. Wirth, A. J.; Gruebele, M. (2013). "Quinary protein structure and the consequences of crowding in living cells: Leaving the test-tube behind". BioEssays. 35 (11): 984–993. doi:10.1002/bies.201300080. PMID   23943406. S2CID   33478753.
  13. Peter B. Crowley; Elysian Chow; Tatiana Papkovskaia (2011). "Protein Interactions in the Escherichia coli Cytosol: An Impediment to In-Cell NMR Spectroscopy". ChemBioChem. 12 (7): 1043–1048. doi:10.1002/cbic.201100063. PMID   21448871. S2CID   44250541.
  14. Xin Mu; Seongil Choi; Lisa Lang; David Mowray; Nikolay V. Dokholyan; Jens Danielsson; Mikael Oliveberg (2017). "Physicochemical code for quinary protein interactions in Escherichia coli". Proceedings of the National Academy of Sciences of the United States of America. 114 (23): E4556–E4563. Bibcode:2017PNAS..114E4556M. doi: 10.1073/pnas.1621227114 . PMC   5468600 . PMID   28536196.