Post-translational modification

Last updated
Post-translational modification of insulin. At the top, the ribosome translates a mRNA sequence into a protein, insulin, and passes the protein through the endoplasmic reticulum, where it is cut, folded, and held in shape by disulfide (-S-S-) bonds. Then the protein passes through the golgi apparatus, where it is packaged into a vesicle. In the vesicle, more parts are cut off, and it turns into mature insulin. Insulin path.svg
Post-translational modification of insulin. At the top, the ribosome translates a mRNA sequence into a protein, insulin, and passes the protein through the endoplasmic reticulum, where it is cut, folded, and held in shape by disulfide (-S-S-) bonds. Then the protein passes through the golgi apparatus, where it is packaged into a vesicle. In the vesicle, more parts are cut off, and it turns into mature insulin.

Post-translational modification (PTM) is the covalent process of changing proteins following protein biosynthesis. PTMs may involve enzymes or occur spontaneously. Proteins are created by ribosomes translating mRNA into polypeptide chains, which may then change to form the mature protein product. PTMs are important components in cell signalling, as for example when prohormones are converted to hormones.

Contents

Post-translational modifications can occur on the amino acid side chains or at the protein's C- or N- termini. [1] They can expand the chemical set of the 22 amino acids by changing an existing functional group or adding a new one such as phosphate. Phosphorylation is highly effective for controlling the enzyme activity and is the most common change after translation. [2] Many eukaryotic and prokaryotic proteins also have carbohydrate molecules attached to them in a process called glycosylation, which can promote protein folding and improve stability as well as serving regulatory functions. Attachment of lipid molecules, known as lipidation, often targets a protein or part of a protein attached to the cell membrane.

Other forms of post-translational modification consist of cleaving peptide bonds, as in processing a propeptide to a mature form or removing the initiator methionine residue. The formation of disulfide bonds from cysteine residues may also be referred to as a post-translational modification. [3] For instance, the peptide hormone insulin is cut twice after disulfide bonds are formed, and a propeptide is removed from the middle of the chain; the resulting protein consists of two polypeptide chains connected by disulfide bonds.

Some types of post-translational modification are consequences of oxidative stress. Carbonylation is one example that targets the modified protein for degradation and can result in the formation of protein aggregates. [4] [5] Specific amino acid modifications can be used as biomarkers indicating oxidative damage. [6]

Sites that often undergo post-translational modification are those that have a functional group that can serve as a nucleophile in the reaction: the hydroxyl groups of serine, threonine, and tyrosine; the amine forms of lysine, arginine, and histidine; the thiolate anion of cysteine; the carboxylates of aspartate and glutamate; and the N- and C-termini. In addition, although the amide of asparagine is a weak nucleophile, it can serve as an attachment point for glycans. Rarer modifications can occur at oxidized methionines and at some methylene groups in side chains. [7]

Post-translational modification of proteins can be experimentally detected by a variety of techniques, including mass spectrometry, Eastern blotting, and Western blotting. Additional methods are provided in the #External links section.

PTMs involving addition of functional groups

Addition by an enzyme in vivo

Hydrophobic groups for membrane localization

Cofactors for enhanced enzymatic activity

Modifications of translation factors

Smaller chemical groups

Non-enzymatic modifications in vivo

Examples of non-enzymatic PTMs are glycation, glycoxidation, nitrosylation, oxidation, succination, and lipoxidation. [15]

Non-enzymatic additions in vitro

Conjugation with other proteins or peptides

Chemical modification of amino acids

Structural changes

Statistics

Common PTMs by frequency

In 2011, statistics of each post-translational modification experimentally and putatively detected have been compiled using proteome-wide information from the Swiss-Prot database. [24] The 10 most common experimentally found modifications were as follows: [25]

FrequencyModification
58383 Phosphorylation
6751 Acetylation
5526 N-linked glycosylation
2844 Amidation
1619 Hydroxylation
1523 Methylation
1133 O-linked glycosylation
878 Ubiquitylation
826 Pyrrolidone carboxylic acid
504 Sulfation

Common PTMs by residue

Some common post-translational modifications to specific amino-acid residues are shown below. Modifications occur on the side-chain unless indicated otherwise.

Amino AcidAbbrev.Modification
Alanine Ala or A N-acetylation (N-terminus)
Arginine Arg or Rdeimination to citrulline, methylation
Asparagine Asn or N deamidation to Asp or iso(Asp), N-linked glycosylation, spontaneous isopeptide bond formation
Aspartic acid Asp or D isomerization to isoaspartic acid, spontaneous isopeptide bond formation
Cysteine Cys or C disulfide-bond formation, oxidation to sulfenic, sulfinic or sulfonic acid, palmitoylation, N-acetylation (N-terminus), S-nitrosylation
Glutamine Gln or Qcyclization to pyroglutamic acid (N-terminus), deamidation to Glutamic acid or isopeptide bond formation to a lysine by a transglutaminase
Glutamic acid Glu or Ecyclization to Pyroglutamic acid (N-terminus), gamma-carboxylation
Glycine Gly or GN-Myristoylation (N-terminus), N-acetylation (N-terminus)
Histidine His or H Phosphorylation
Isoleucine Ile or I
Leucine Leu or L
Lysine Lys or K acetylation, ubiquitylation, SUMOylation, methylation, hydroxylation leading to allysine, spontaneous isopeptide bond formation
Methionine Met or M N-acetylation (N-terminus), N-linked Ubiquitination, oxidation to sulfoxide or sulfone
Phenylalanine Phe or F
Proline Pro or P hydroxylation
Serine Ser or S Phosphorylation, O-linked glycosylation, N-acetylation (N-terminus)
Threonine Thr or T Phosphorylation, O-linked glycosylation, N-acetylation (N-terminus)
Tryptophan Trp or Wmono- or di-oxidation, formation of kynurenine, tryptophan tryptophylquinone
Tyrosine Tyr or Y sulfation, phosphorylation
Valine Val or V N-acetylation (N-terminus)

Databases and tools

Flowchart of the process and the data sources to predict PTMs. Image for Wiki 2.jpg
Flowchart of the process and the data sources to predict PTMs.

Protein sequences contain sequence motifs that are recognized by modifying enzymes, and which can be documented or predicted in PTM databases. With the large number of different modifications being discovered, there is a need to document this sort of information in databases. PTM information can be collected through experimental means or predicted from high-quality, manually curated data. Numerous databases have been created, often with a focus on certain taxonomic groups (e.g. human proteins) or other features.

List of resources

Tools

List of software for visualization of proteins and their PTMs

Case examples

See also

Related Research Articles

<span class="mw-page-title-main">Amino acid</span> Organic compounds containing amine and carboxylic groups

Amino acids are organic compounds that contain both amino and carboxylic acid functional groups. Although over 500 amino acids exist in nature, by far the most important are the 22 α-amino acids incorporated into proteins. Only these 22 appear in the genetic code of life.

<span class="mw-page-title-main">Protein primary structure</span> Linear sequence of amino acids in a peptide or protein

Protein primary structure is the linear sequence of amino acids in a peptide or protein. By convention, the primary structure of a protein is reported starting from the amino-terminal (N) end to the carboxyl-terminal (C) end. Protein biosynthesis is most commonly performed by ribosomes in cells. Peptides can also be synthesized in the laboratory. Protein primary structures can be directly sequenced, or inferred from DNA sequences.

<span class="mw-page-title-main">Protein biosynthesis</span> Assembly of proteins inside biological cells

Protein biosynthesis is a core biological process, occurring inside cells, balancing the loss of cellular proteins through the production of new proteins. Proteins perform a number of critical functions as enzymes, structural proteins or hormones. Protein synthesis is a very similar process for both prokaryotes and eukaryotes but there are some distinct differences.

<span class="mw-page-title-main">Glycoprotein</span> Protein with oligosaccharide modifications

Glycoproteins are proteins which contain oligosaccharide chains covalently attached to amino acid side-chains. The carbohydrate is attached to the protein in a cotranslational or posttranslational modification. This process is known as glycosylation. Secreted extracellular proteins are often glycosylated.

<span class="mw-page-title-main">Lipid-anchored protein</span> Membrane protein

Lipid-anchored proteins are proteins located on the surface of the cell membrane that are covalently attached to lipids embedded within the cell membrane. These proteins insert and assume a place in the bilayer structure of the membrane alongside the similar fatty acid tails. The lipid-anchored protein can be located on either side of the cell membrane. Thus, the lipid serves to anchor the protein to the cell membrane. They are a type of proteolipids.

Glycosylation is the reaction in which a carbohydrate, i.e. a glycosyl donor, is attached to a hydroxyl or other functional group of another molecule in order to form a glycoconjugate. In biology, glycosylation usually refers to an enzyme-catalysed reaction, whereas glycation may refer to a non-enzymatic reaction.

<span class="mw-page-title-main">Hemagglutinin esterase</span> Glycoprotein present in some enveloped viruses

Hemagglutinin esterase (HEs) is a glycoprotein that certain enveloped viruses possess and use as an invading mechanism. HEs helps in the attachment and destruction of certain sialic acid receptors that are found on the host cell surface. Viruses that possess HEs include influenza C virus, toroviruses, and coronaviruses of the subgenus Embecovirus. HEs is a dimer transmembrane protein consisting of two monomers, each monomer is made of three domains. The three domains are: membrane fusion, esterase, and receptor binding domains.

<span class="mw-page-title-main">Catalytic triad</span> Set of three coordinated amino acids

A catalytic triad is a set of three coordinated amino acids that can be found in the active site of some enzymes. Catalytic triads are most commonly found in hydrolase and transferase enzymes. An acid-base-nucleophile triad is a common motif for generating a nucleophilic residue for covalent catalysis. The residues form a charge-relay network to polarise and activate the nucleophile, which attacks the substrate, forming a covalent intermediate which is then hydrolysed to release the product and regenerate free enzyme. The nucleophile is most commonly a serine or cysteine amino acid, but occasionally threonine or even selenocysteine. The 3D structure of the enzyme brings together the triad residues in a precise orientation, even though they may be far apart in the sequence.

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

Myristoylation is a lipidation modification where a myristoyl group, derived from myristic acid, is covalently attached by an amide bond to the alpha-amino group of an N-terminal glycine residue. Myristic acid is a 14-carbon saturated fatty acid (14:0) with the systematic name of n-tetradecanoic acid. This modification can be added either co-translationally or post-translationally. N-myristoyltransferase (NMT) catalyzes the myristic acid addition reaction in the cytoplasm of cells. This lipidation event is the most common type of fatty acylation and is present in many organisms, including animals, plants, fungi, protozoans and viruses. Myristoylation allows for weak protein–protein and protein–lipid interactions and plays an essential role in membrane targeting, protein–protein interactions and functions widely in a variety of signal transduction pathways.

<span class="mw-page-title-main">Isopeptide bond</span>

An isopeptide bond is a type of amide bond formed between a carboxyl group of one amino acid and an amino group of another. An isopeptide bond is the linkage between the side chain amino or carboxyl group of one amino acid to the α-carboxyl, α-amino group, or the side chain of another amino acid. In a typical peptide bond, also known as eupeptide bond, the amide bond always forms between the α-carboxyl group of one amino acid and the α-amino group of the second amino acid. Isopeptide bonds are rarer than regular peptide bonds. Isopeptide bonds lead to branching in the primary sequence of a protein. Proteins formed from normal peptide bonds typically have a linear primary sequence.

Bioconjugation is a chemical strategy to form a stable covalent link between two molecules, at least one of which is a biomolecule.

<span class="mw-page-title-main">Histone-modifying enzymes</span> Type of enzymes

Histone-modifying enzymes are enzymes involved in the modification of histone substrates after protein translation and affect cellular processes including gene expression. To safely store the eukaryotic genome, DNA is wrapped around four core histone proteins, which then join to form nucleosomes. These nucleosomes further fold together into highly condensed chromatin, which renders the organism's genetic material far less accessible to the factors required for gene transcription, DNA replication, recombination and repair. Subsequently, eukaryotic organisms have developed intricate mechanisms to overcome this repressive barrier imposed by the chromatin through histone modification, a type of post-translational modification which typically involves covalently attaching certain groups to histone residues. Once added to the histone, these groups elicit either a loose and open histone conformation, euchromatin, or a tight and closed histone conformation, heterochromatin. Euchromatin marks active transcription and gene expression, as the light packing of histones in this way allows entry for proteins involved in the transcription process. As such, the tightly packed heterochromatin marks the absence of current gene expression.

<span class="mw-page-title-main">ADP-ribosylation</span> Addition of one or more ADP-ribose moieties to a protein.

ADP-ribosylation is the addition of one or more ADP-ribose moieties to a protein. It is a reversible post-translational modification that is involved in many cellular processes, including cell signaling, DNA repair, gene regulation and apoptosis. Improper ADP-ribosylation has been implicated in some forms of cancer. It is also the basis for the toxicity of bacterial compounds such as cholera toxin, diphtheria toxin, and others.

<span class="mw-page-title-main">Protein phosphorylation</span> Process of introducing a phosphate group on to a protein

Protein phosphorylation is a reversible post-translational modification of proteins in which an amino acid residue is phosphorylated by a protein kinase by the addition of a covalently bound phosphate group. Phosphorylation alters the structural conformation of a protein, causing it to become activated, deactivated, or otherwise modifying its function. Approximately 13,000 human proteins have sites that are phosphorylated.

Glycopeptides are peptides that contain carbohydrate moieties (glycans) covalently attached to the side chains of the amino acid residues that constitute the peptide.

O-linked glycosylation is the attachment of a sugar molecule to the oxygen atom of serine (Ser) or threonine (Thr) residues in a protein. O-glycosylation is a post-translational modification that occurs after the protein has been synthesised. In eukaryotes, it occurs in the endoplasmic reticulum, Golgi apparatus and occasionally in the cytoplasm; in prokaryotes, it occurs in the cytoplasm. Several different sugars can be added to the serine or threonine, and they affect the protein in different ways by changing protein stability and regulating protein activity. O-glycans, which are the sugars added to the serine or threonine, have numerous functions throughout the body, including trafficking of cells in the immune system, allowing recognition of foreign material, controlling cell metabolism and providing cartilage and tendon flexibility. Because of the many functions they have, changes in O-glycosylation are important in many diseases including cancer, diabetes and Alzheimer's. O-glycosylation occurs in all domains of life, including eukaryotes, archaea and a number of pathogenic bacteria including Burkholderia cenocepacia, Neisseria gonorrhoeae and Acinetobacter baumannii.

<span class="mw-page-title-main">Non-proteinogenic amino acids</span> Are not naturally encoded in the genome

In biochemistry, non-coded or non-proteinogenic amino acids are distinct from the 22 proteinogenic amino acids which are naturally encoded in the genome of organisms for the assembly of proteins. However, over 140 non-proteinogenic amino acids occur naturally in proteins and thousands more may occur in nature or be synthesized in the laboratory. Chemically synthesized amino acids can be called unnatural amino acids. Unnatural amino acids can be synthetically prepared from their native analogs via modifications such as amine alkylation, side chain substitution, structural bond extension cyclization, and isosteric replacements within the amino acid backbone. Many non-proteinogenic amino acids are important:

Protein <i>O</i>-GlcNAc transferase Protein-coding gene in the species Homo sapiens

Protein O-GlcNAc transferase also known as OGT or O-linked N-acetylglucosaminyltransferase is an enzyme that in humans is encoded by the OGT gene. OGT catalyzes the addition of the O-GlcNAc post-translational modification to proteins.

Protein methylation is a type of post-translational modification featuring the addition of methyl groups to proteins. It can occur on the nitrogen-containing side-chains of arginine and lysine, but also at the amino- and carboxy-termini of a number of different proteins. In biology, methyltransferases catalyze the methylation process, activated primarily by S-adenosylmethionine. Protein methylation has been most studied in histones, where the transfer of methyl groups from S-adenosyl methionine is catalyzed by histone methyltransferases. Histones that are methylated on certain residues can act epigenetically to repress or activate gene expression.

<span class="mw-page-title-main">Dehydroamino acid</span>

In biochemistry, a dehydroamino acid or α,β-dehydroamino acid is an amino acids, usually with a C=C double bond in its side chain. Dehydroamino acids are not coded by DNA, but arise via post-translational modification.

References

  1. Pratt, Charlotte W.; Voet, Judith G.; Voet, Donald (2006). Fundamentals of Biochemistry: Life at the Molecular Level (2nd ed.). Hoboken, NJ: Wiley. ISBN   9780471214953. OCLC   1280801548. Archived from the original on 13 July 2012.
  2. Khoury GA, Baliban RC, Floudas CA (September 2011). "Proteome-wide post-translational modification statistics: frequency analysis and curation of the swiss-prot database". Scientific Reports . 1: 90. Bibcode:2011NatSR...1E..90K. doi:10.1038/srep00090. PMC   3201773 . PMID   22034591.
  3. Lodish H, Berk A, Zipursky SL, et al. (2000). "17.6, Post-Translational Modifications and Quality Control in the Rough ER". Molecular Cell Biology (4th ed.). New York: W. H. Freeman. ISBN   978-0-7167-3136-8.
  4. Dalle-Donne I, Aldini G, Carini M, Colombo R, Rossi R, Milzani A (2006). "Protein carbonylation, cellular dysfunction, and disease progression". Journal of Cellular and Molecular Medicine . 10 (2): 389–406. doi:10.1111/j.1582-4934.2006.tb00407.x. PMC   3933129 . PMID   16796807.
  5. Grimsrud PA, Xie H, Griffin TJ, Bernlohr DA (August 2008). "Oxidative stress and covalent modification of protein with bioactive aldehydes". The Journal of Biological Chemistry . 283 (32): 21837–41. doi: 10.1074/jbc.R700019200 . PMC   2494933 . PMID   18445586.
  6. Gianazza E, Crawford J, Miller I (July 2007). "Detecting oxidative post-translational modifications in proteins". Amino Acids. 33 (1): 51–6. doi:10.1007/s00726-006-0410-2. PMID   17021655. S2CID   23819101.
  7. Walsh, Christopher T. (2006). Posttranslational modification of proteins : expanding nature's inventory. Englewood: Roberts and Co. Publ. ISBN   9780974707730.:12–14
  8. Whiteheart SW, Shenbagamurthi P, Chen L, Cotter RJ, Hart GW, et al. (August 1989). "Murine elongation factor 1 alpha (EF-1 alpha) is posttranslationally modified by novel amide-linked ethanolamine-phosphoglycerol moieties. Addition of ethanolamine-phosphoglycerol to specific glutamic acid residues on EF-1 alpha". The Journal of Biological Chemistry. 264 (24): 14334–41. doi: 10.1016/S0021-9258(18)71682-7 . PMID   2569467.
  9. Roy H, Zou SB, Bullwinkle TJ, Wolfe BS, Gilreath MS, Forsyth CJ, Navarre WW, Ibba M (August 2011). "The tRNA synthetase paralog PoxA modifies elongation factor-P with (R)-β-lysine". Nature Chemical Biology. 7 (10): 667–9. doi:10.1038/nchembio.632. PMC   3177975 . PMID   21841797.
  10. Ali I, Conrad RJ, Verdin E, Ott M (February 2018). "Lysine Acetylation Goes Global: From Epigenetics to Metabolism and Therapeutics". Chem Rev. 118 (3): 1216–1252. doi:10.1021/acs.chemrev.7b00181. PMC   6609103 . PMID   29405707.
  11. Bradbury AF, Smyth DG (March 1991). "Peptide amidation". Trends in Biochemical Sciences. 16 (3): 112–5. doi:10.1016/0968-0004(91)90044-v. PMID   2057999.
  12. Eddé B, Rossier J, Le Caer JP, Desbruyères E, Gros F, Denoulet P (January 1990). "Posttranslational glutamylation of alpha-tubulin". Science. 247 (4938): 83–5. Bibcode:1990Sci...247...83E. doi:10.1126/science.1967194. PMID   1967194.
  13. Walker CS, Shetty RP, Clark K, Kazuko SG, Letsou A, Olivera BM, Bandyopadhyay PK, et al. (March 2001). "On a potential global role for vitamin K-dependent gamma-carboxylation in animal systems. Evidence for a gamma-glutamyl carboxylase in Drosophila". The Journal of Biological Chemistry. 276 (11): 7769–74. doi: 10.1074/jbc.M009576200 . PMID   11110799.
  14. 1 2 3 Chung HS, et al. (January 2013). "Cysteine oxidative posttranslational modifications: emerging regulation in the cardiovascular system". Circulation Research. 112 (2): 382–92. doi:10.1161/CIRCRESAHA.112.268680. PMC   4340704 . PMID   23329793.
  15. "The Advanced Lipoxidation End-Product Malondialdehyde-Lysine in Aging and Longevity" PMID 33203089 PMC7696601
  16. Jaisson S, Pietrement C, Gillery P (November 2011). "Carbamylation-derived products: bioactive compounds and potential biomarkers in chronic renal failure and atherosclerosis". Clinical Chemistry. 57 (11): 1499–505. doi: 10.1373/clinchem.2011.163188 . PMID   21768218.
  17. Kang HJ, Baker EN (April 2011). "Intramolecular isopeptide bonds: protein crosslinks built for stress?". Trends in Biochemical Sciences. 36 (4): 229–37. doi:10.1016/j.tibs.2010.09.007. PMID   21055949.
  18. Stark GR, Stein WH, Moore X (1960). "Reactions of the Cyanate Present in Aqueous Urea with Amino Acids and Proteins". J Biol Chem. 235 (11): 3177–3181. doi: 10.1016/S0021-9258(20)81332-5 .
  19. Van G. Wilson (Ed.) (2004). Sumoylation: Molecular Biology and Biochemistry Archived 2005-02-09 at the Wayback Machine . Horizon Bioscience. ISBN   0-9545232-8-8.
  20. Malakhova OA, Yan M, Malakhov MP, Yuan Y, Ritchie KJ, Kim KI, Peterson LF, Shuai K, Zhang DE (February 2003). "Protein ISGylation modulates the JAK-STAT signaling pathway". Genes & Development. 17 (4): 455–60. doi:10.1101/gad.1056303. PMC   195994 . PMID   12600939.
  21. Klareskog L, Rönnelid J, Lundberg K, Padyukov L, Alfredsson L (2008). "Immunity to citrullinated proteins in rheumatoid arthritis". Annual Review of Immunology. 26: 651–75. doi:10.1146/annurev.immunol.26.021607.090244. PMID   18173373.
  22. Brennan DF, Barford D (March 2009). "Eliminylation: a post-translational modification catalyzed by phosphothreonine lyases". Trends in Biochemical Sciences. 34 (3): 108–14. doi:10.1016/j.tibs.2008.11.005. PMID   19233656.
  23. Rabe von Pappenheim, Fabian; Wensien, Marie; Ye, Jin; Uranga, Jon; Irisarri, Iker; de Vries, Jan; Funk, Lisa-Marie; Mata, Ricardo A.; Tittmann, Kai (April 2022). "Widespread occurrence of covalent lysine–cysteine redox switches in proteins". Nature Chemical Biology. 18 (4): 368–375. doi: 10.1038/s41589-021-00966-5 .
  24. Khoury GA, Baliban RC, Floudas CA (September 2011). "Proteome-wide post-translational modification statistics: frequency analysis and curation of the swiss-prot database". Scientific Reports. 1 (90): 90. Bibcode:2011NatSR...1E..90K. doi:10.1038/srep00090. PMC   3201773 . PMID   22034591.
  25. "Proteome-Wide Post-Translational Modification Statistics". selene.princeton.edu. Archived from the original on 2012-08-30. Retrieved 2011-07-22.
  26. 1 2 Lee TY, Huang HD, Hung JH, Huang HY, Yang YS, Wang TH (January 2006). "dbPTM: an information repository of protein post-translational modification". Nucleic Acids Research. 34 (Database issue): D622-7. doi:10.1093/nar/gkj083. PMC   1347446 . PMID   16381945.
  27. Hornbeck PV, Zhang B, Murray B, Kornhauser JM, Latham V, Skrzypek E (January 2015). "PhosphoSitePlus, 2014: mutations, PTMs and recalibrations". Nucleic Acids Research. 43 (Database issue): D512-20. doi:10.1093/nar/gku1267. PMC   4383998 . PMID   25514926.
  28. 1 2 Goel R, Harsha HC, Pandey A, Prasad TS (February 2012). "Human Protein Reference Database and Human Proteinpedia as resources for phosphoproteome analysis". Molecular BioSystems. 8 (2): 453–63. doi:10.1039/c1mb05340j. PMC   3804167 . PMID   22159132.
  29. Sigrist CJ, Cerutti L, de Castro E, Langendijk-Genevaux PS, Bulliard V, Bairoch A, Hulo N (January 2010). "PROSITE, a protein domain database for functional characterization and annotation". Nucleic Acids Research. 38 (Database issue): D161-6. doi:10.1093/nar/gkp885. PMC   2808866 . PMID   19858104.
  30. Garavelli JS (January 2003). "The RESID Database of Protein Modifications: 2003 developments". Nucleic Acids Research. 31 (1): 499–501. doi:10.1093/nar/gkg038. PMC   165485 . PMID   12520062.
  31. Huang H, Arighi CN, Ross KE, Ren J, Li G, Chen SC, Wang Q, Cowart J, Vijay-Shanker K, Wu CH (January 2018). "iPTMnet: an integrated resource for protein post-translational modification network discovery". Nucleic Acids Research. 46 (1): D542–D550. doi:10.1093/nar/gkx1104. PMC   5753337 . PMID   2914561.
  32. Audagnotto M, Dal Peraro M (2017-03-31). "In silico prediction tools and molecular modeling". Computational and Structural Biotechnology Journal. 15: 307–319. doi:10.1016/j.csbj.2017.03.004. PMC   5397102 . PMID   28458782.
  33. Wulff-Fuentes E, Berendt RR, Massman L, Danner L, Malard F, Vora J, Kahsay R, Olivier-Van Stichelen S (January 2021). "The human O-GlcNAcome database and meta-analysis". Scientific Data. 8 (1): 25. Bibcode:2021NatSD...8...25W. doi:10.1038/s41597-021-00810-4. PMC   7820439 . PMID   33479245.
  34. Malard F, Wulff-Fuentes E, Berendt RR, Didier G, Olivier-Van Stichelen S (July 2021). "Automatization and self-maintenance of the O-GlcNAcome catalog: a smart scientific database". Database (Oxford). 2021: 1. doi:10.1093/database/baab039. PMC   8288053 . PMID   34279596.
  35. Warnecke A, Sandalova T, Achour A, Harris RA (November 2014). "PyTMs: a useful PyMOL plugin for modeling common post-translational modifications". BMC Bioinformatics. 15 (1): 370. doi: 10.1186/s12859-014-0370-6 . PMC   4256751 . PMID   25431162.
  36. Yang Y, Peng X, Ying P, Tian J, Li J, Ke J, Zhu Y, Gong Y, Zou D, Yang N, Wang X, Mei S, Zhong R, Gong J, Chang J, Miao X (January 2019). "AWESOME: a database of SNPs that affect protein post-translational modifications". Nucleic Acids Research. 47 (D1): D874–D880. doi:10.1093/nar/gky821. PMC   6324025 . PMID   30215764.
  37. Morris JH, Huang CC, Babbitt PC, Ferrin TE (September 2007). "structureViz: linking Cytoscape and UCSF Chimera". Bioinformatics. 23 (17): 2345–7. doi: 10.1093/bioinformatics/btm329 . PMID   17623706.
  38. "1tp8 - Proteopedia, life in 3D". www.proteopedia.org.

(Wayback Machine copy)

(Wayback Machine copy)