Self-hydroxylation of the splicing factor lysyl hydroxylase, JMJD6

Monica Mantri a, Celia J. Webby a, Nikita D. Loik a, Refaat B. Hamed ab, Michael L. Nielsen c, Michael A. McDonough a, James S. O. McCullagh a, Angelika Böttger d, Christopher J. Schofield *a and Alexander Wolf *a
aChemistry Research Laboratory, 12 Mansfield Rd., Oxford, OX1 3TA, United Kingdom. E-mail: christopher.schofield@chem.ox.ac.uk; alexander.wolf@chem.ox.ac.uk; Fax: +44 (0)1865 285022
bDepartment of Pharmacognosy, Faculty of Pharmacy, Assiut University, 71526, Egypt
cDepartment of Proteomics, The Novo Nordisk Foundation Center for Protein Research, University of Copenhagen, Faculty of Health Sciences, DK-2200, Copenhagen, Denmark
dDepartment of Biology II, Ludwig-Maximilians-University, Munich, Großhaderner Str 2, D-82152, Planegg-Martinsried, Germany

Received 1st September 2011 , Accepted 19th October 2011

First published on 18th November 2011


Abstract

The lysyl5S-hydroxylase, JMJD6 acts on proteins involved in RNA splicing. We find that in the absence of substrate JMJD6 catalyses turnover of 2OG to COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
succinate
. 1H-NMR analyses demonstrate that consumption of 2OG is coupled to COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
succinate
formation. MS analyses reveal that JMJD6 undergoes self-hydroxylation in the presence of COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and 2OG resulting in production of 5S-hydroxylysine residues. JMJD6 in human cells is also found to be hydroxylated. Self-hydroxylation of JMJD6 may play a regulatory role in modulating the hydroxylation status of proteins involved in RNA splicing.


JMJD6 has been reported as a lysyl5S-hydroxylase active on proteins involved in RNA splicing.1–3JMJD6 has also been reported to be a histone methyl arginine demethylase,4 though we have not yet observed this activity under our assay conditions.1,2JMJD6 belongs to the COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
2-oxoglutarate
(2OG) and COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
dependent family of oxygenases which catalyses a diverse range of oxidative reactions.5,6 Various human 2OG oxygenases are being explored as therapeutic targets via small molecule inhibition, including the hypoxia inducible factor (HIF) hydroxylases and the JmjC domain containing Nε-methyl lysine histone demethylases.7 The JmjC enzymes are of particular interest because of their important role in epigenetics. The 2OG oxygenases utilise 2OG as a co-substrate and share a conserved COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
binding motif. In addition to coupling 2OGoxidation to that of substrate, many 2OG oxygenases catalyse the uncoupled turnover of 2OG, i.e. where 2OG is converted to COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
succinate
in the absence of a primary substrate, albeit typically at a substantially lower rate than substrate coupled turnover.8–12 Several members of the 2OG oxygenase family have also been shown to catalyse self-oxidation in the absence of substrate. Most of the reported self-hydroxylation reactions to date involve hydroxylation of an active site aromatic residue located within ∼10 Å of the active site COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
iron
.9–11

The closest human homologue of JMJD6 is Factor Inhibiting HIF (FIH),13 which reduces the transcriptional activity of HIF by post translational asparaginylhydroxylation in the C-terminal transcriptional activation domain of HIF-α.5FIH has been reported to undergo self-hydroxylation of an active site tryptophan residue9 in the absence of HIF. Other members of the 2OG oxygenase family, taurine dioxygenase (TauD) and 2,4-dichlorophenoxyacetate oxygenase (TfdA) are apparently deactivated by hydroxylation of an active site tyrosine (TauD)10 and tryptophan (TfdA)11 residues in the absence of their primary substrates, COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
taurine
or 2,4-dichlorophenoxyacetate, respectively. In the case of the human AlkBhomologhABH3, there is evidence that a leucine residue (Leu177hABH3) undergoes oxidation to give an alcohol which can be further oxidised to an aldehyde at the leucine δ-position.14 Oxidative cleavage has also been reported in case of COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
1-aminocyclopropane-1-carboxylate
(ACC) oxidase, the COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
ethylene
forming enzyme in plants which is closely related to 2OG oxygenases.15

It has been speculated that the self-regulation of some 2OG oxygenases may protect cells from damaging COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
oxygen
derived species or may serve as a negative feedback mechanism.9 Self-regulation is employed by many enzymes in nature, e.g. kinases autophosphorylate residues on their own protein scaffold to regulate their activity16 and ubiquitinases are capable of targeting themselves for ubiquitin-dependent degradation.17,18

Here we demonstrate self-hydroxylation of the JMJD6protein. Unlike some other members of the 2OG oxygenase family that undergo self-hydroxylation involving oxidation of aromatic/active site residues, recombinant JMJD6hydroxylates lysine residues (Lys111JMJD6 and Lys167JMJD6) in the same manner as its splicing protein substrates.1 We further report that endogenous JMJD6 extracted from human cells undergoes self-hydroxylation.

Results

To investigate the uncoupled turnover of 2OG by JMJD6 we began by carrying out assays measuring the conversion of [1-14C]-2OG into 14CO219 in the presence and the absence of the human peptide substrate LUC7L2267–278, which is a fragment of splicing regulatory protein LUC7like2.1 Under standard assay conditions we found that the amount of 2OG consumed was 1.7 fold higher in the presence of the LUC7L2267–278peptide than in its absence (Fig. 1A and Fig. 3 in ref. 29), suggesting that uncoupled turnover by JMJD6 might be relatively high even in the presence of a hydroxylated substrate (albeit only a fraction of a protein). 1H-NMR analyses verified the uncoupled turnover of 2OG to COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
succinate
in the absence of splicing proteinpeptide substrate (Fig. 1B) in an COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and 2OG dependent manner. We then carried out investigations using a MALDI-MS based assay that monitors substrate hydroxylation by a +16Da mass shift related to the unhydroxylated substrate peptide. The results suggest that when JMJD6 was pre-incubated with 2OG, COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
for 30 min prior to the addition of LUC7L2267–278, a 40–50% decrease in substrate hydroxylation was observed in comparison to when all the components were mixed simultaneously (results not shown). In some but not all cases, coupled turnover of 2OG is proportional to the oxidation of ascorbatee.g. the collagenprolyl-4-hydroxylases20–22 and references therein. However, in our work the JMJD6 incubations were carried out in the absence of ascorbate. Although the relatively high level of uncoupled turnover in the presence of substrate may reflect use of a non optimal substrate, we considered that JMJD6 might accomplish (some) catalyticoxidationvia a self-hydroxylation process.

Radioactive assays, 1H-NMR and amino acid analysis showing 2OG turnover catalysed by JMJD6. (A) Results of [1-14C]-turnover assay under standard assay conditions with Fe(ii), 2OG and LUC7L2peptide, Values are Mean ± S.D., n = 3; (B) JMJD6 catalysed 2OG turnover in the absence of substrate as observed by 1H-NMR; (C) and (D) Amino acid analyses of JMJD6 catalysis. After incubation, the peptide product was hydrolysed using trypsin and carboxypeptidase Y. (Ci) Hydroxylsyine product obtained after hydrolysis of LUC7L2267–278peptide that had been incubated with JMJD6, Fe(ii) and 2OG; (Cii) Hydroxylsyine product obtained after hydrolysis of JMJD6 that had been incubated with Fe(ii) and 2OG; (Ciii) 2S,5R-Hydroxylysine standard; (Civ) Mixture of 2R,5R/2S,5S- and 2R,5S/2S,5R-hydroxylysine standards; (Di) Hydroxylysine product from JMJD6 as purified after hydrolysis, i.e. without Fe(ii) and 2OG incubation; (Dii) Hydroxylysine product obtained after incubation of JMJD6 incubated with Fe(ii); (Diii) Hydrolysis product obtained after incubation of JMJD6 with 2OG; (Div) Hydroxylysine product obtained after incubation of JMJD6 with Fe(ii) and 2OG.
Fig. 1 Radioactive assays, 1H-NMR and amino acid analysis showing 2OG turnover catalysed by JMJD6. (A) Results of [1-14C]-turnover assay under standard assay conditions with COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
, 2OG and LUC7L2peptide, Values are Mean ± S.D., n = 3; (B) JMJD6 catalysed 2OG turnover in the absence of substrate as observed by 1H-NMR; (C) and (D) Amino acid analyses of JMJD6 catalysis. After incubation, the peptide product was hydrolysed using trypsin and carboxypeptidase Y. (Ci) Hydroxylsyine product obtained after hydrolysis of LUC7L2267–278peptide that had been incubated with JMJD6, COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and 2OG; (Cii) Hydroxylsyine product obtained after hydrolysis of JMJD6 that had been incubated with COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and 2OG; (Ciii) 2S,5R-Hydroxylysine standard; (Civ) Mixture of 2R,5R/2S,5S- and 2R,5S/2S,5R-hydroxylysine standards; (Di) COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Hydroxylysine
product from JMJD6 as purified after hydrolysis, i.e. without COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and 2OG incubation; (Dii) COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Hydroxylysine
product obtained after incubation of JMJD6 incubated with COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
; (Diii) Hydrolysis product obtained after incubation of JMJD6 with 2OG; (Div) COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Hydroxylysine
product obtained after incubation of JMJD6 with COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and 2OG.

To test for self-hydroxylation of JMJD6, we first carried out amino acid analyses on recombinant JMJD6 that had been incubated with or without COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and/or 2OG. The JMJD6hydrolysis procedure employed was the one which we had previously reported for splicing regulatory peptideLUC7L2267–278 product analysis, i.e. trypsin and carboxypeptidase Y were used rather than acid hydrolysis in order to avoid isomerisation of the COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
hydroxylysine
product.2 The results reveal that JMJD6 undergoes self-hydroxylation on one or more lysine residues in a manner that is stimulated by 2OG. Notably, we found that as purified, recombinant JMJD6 had already undergone a degree of self-hydroxylation presumably in Escherichia colicells (Fig. 1D). The extent of hydroxylation was not substantially increased by incubation with COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
(Fig. 1Dii), likely reflecting co-purification of JMJD6 with COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
. However, the extent of hydroxylation was increased substantially by addition of 2OG (Fig. 1Diii). Importantly, the stereochemistry of the COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
hydroxylysine
product was found to be 2S,5S, i.e. the same as that observed for JMJD6 catalysed LUC7L2hydroxylation (Fig. 1C). The lack of evidence for 5R-lysylhydroxylation as shown in Scheme 1 supports the proposal that the self-hydroxylation of lysines in JMJD6 occurs at the active site in a manner similar to splicing protein hydroxylation.



          JMJD6 catalysed self-hydroxylation gives the 2S,5S-hydroxylysine product as observed for hydroxylation of splicing regulatory protein substrates.
Scheme 1 JMJD6 catalysed self-hydroxylation gives the 2S,5S-hydroxylysine product as observed for hydroxylation of splicing regulatory protein substrates.

We then investigated the site(s) of lysinehydroxylation in recombinant JMJD6 that had been incubated with COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and 2OG using LysCdigestion followed by LCMS/MS analyses.23 These analyses support hydroxylation of Lys111JMJD6 and Lys167JMJD6 (Fig. 2). No hydroxylation at Lys167JMJD6 was observed in the absence of 2OG and COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
(Fig. S1). Hydroxylation at Lys167JMJD6 was also observed to occur in the presence of recombinant U2AF65 substrate, when incubated with COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and 2OG (Fig. S2). We cannot rule out the possibility of hydroxylations at other sites because the proteomic analyses were limited (at least for confident assignment) to fragments containing residues Lys91, Lys100, Lys111, Lys113, Lys115, Lys142, Lys151, Lys154, Lys167, Lys219, Lys317 and Lys397. Indeed the observation that COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
hydroxylysine
is present in recombinant JMJD6 as determined by amino acid analysis reveals that hydroxylation at other sites is likely (Fig. 1D).



          MS/MS
          spectra showing hydroxylation of recombinant JMJD6 in the presence of 2OG and Fe(ii). LC-MS/MS analyses supports hydroxylation at Lys111JMJD6 (a′) and Lys167JMJD6 (b′) of recombinant JMJD6 in the presence of Fe(ii) and 2OG. Hydroxylation was not observed to be complete and the spectra corresponding to unhydroxylated Lys111JMJD6 (a) and Lys167JMJD6 (b) are also shown. The MH2+peptide that was fragmented to give the resulting MSspectra is shown on the right hand side of each spectrum. The fragmented peptide sequence of resulting MS is shown on right hand side and MH2+ precursor ion is shown on left hand side of each spectra.
Fig. 2 MS/MS spectra showing hydroxylation of recombinant JMJD6 in the presence of 2OG and COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
. LC-MS/MS analyses supports hydroxylation at Lys111JMJD6 (a′) and Lys167JMJD6 (b′) of recombinant JMJD6 in the presence of COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
Fe(II)
and 2OG. Hydroxylation was not observed to be complete and the spectra corresponding to unhydroxylated Lys111JMJD6 (a) and Lys167JMJD6 (b) are also shown. The MH2+peptide that was fragmented to give the resulting MSspectra is shown on the right hand side of each spectrum. The fragmented peptide sequence of resulting MS is shown on right hand side and MH2+ precursor ion is shown on left hand side of each spectra.

We then made a set of peptides as shown in Table 1 containing the identified hydroxylation sites at Lys111JMJD6 (with recombinant protein) and Lys167JMJD6 (with recombinant and endogenous protein). Out of the 10 peptides analysed, only two were substrates as observed by MALDI-MS analysis, corresponding to JMJD61–14 and JMJD6301–314. Notably the two peptides, JMJD6105–120, JMJD6167–176 containing the sites of hydroxylation identified in JMJD6protein (Lys111JMJD6 and Lys167JMJD6) were not observed to be substrates (Fig. S4). Although we cannot be certain, it is possible that the hydroxylation of lysines corresponding to JMJD61–14 and JMJD6301–314 observed in peptide work, may reflect the hydroxylation of purified recombinant JMJD6 as obtained by amino acid analysis.

Table 1 Results of MALDI-TOF analyses of peptides synthesised containing possible lysinehydroxylation sites within JMJD6
No. JMJD6 Sequence Peptide sequence Molecular weight +16 Da Shift in MS analysis
1 JMJD6 1–14 MNHKSKKRIREAKR 1782.15 YES
2 JMJD6 14–30 RSARPELKDSLDWTRHN 2081.27 COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
NO
3 JMJD6 78–95 GWSAQEKWTLERLKRKYR 2335.69 COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
NO
4 JMJD6 86–104 TLERLKRKYRNQKFKCGED 2412.80 COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
NO
5 JMJD6 105–120 NDGYSVKMKMKYYIEY 2032.36 COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
NO
Hydroxylation at K111 observed in protein.
6 JMJD6 135–150 SSYGEHPKRRKLLEDY 1978.19 COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
NO
7 JMJD6 160–176 LFQYAGEKRRPPYRWFV 2213.57 COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
NO
Hydroxylation at K167 observed in protein.
8 JMJD6 301–314 TVRGRPKLSRKWYR 1803.14 YES
9 JMJD6 363–379 SGSEGDGTVHRRKKRRT 1921.15 COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
NO
10 JMJD6 363–381 SGSEGDGTVHRRKKRRTCS 2116.07 COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
NO


Finally, we investigated whether self-hydroxylation of endogenous JMJD6 in human cells occurs. JMJD6 was immunoprecipitated from HeLa cells using an anti-JMJD6 antibody and was isolated by SDS-PAGE. The band corresponding to the expected molecular weight of JMJD6 was excised from an acrylamide gel, digested with LysCprotease and analysed by LC-MS/MS. The fragment ion series was consistent with hydroxylation of Lys167JMJD6 (Fig. 3). As mentioned before we cannot rule out the possibility of hydroxylation at other sites.



          MS/MS analysis of endogenous JMJD6 supports hydroxylation of Lys167JMJD6. JMJD6 was immunoprecipitated from HeLa cells, digested with LysCprotease and analysed by LC-MS/MS. The MH2+peptide that was fragmented to give the resulting MSspectra is shown on the right hand side of the spectrum (in green). The co-elution of two different peptides is shown by use of blue and red colours.
Fig. 3 MS/MS analysis of endogenous JMJD6 supports hydroxylation of Lys167JMJD6. JMJD6 was immunoprecipitated from HeLa cells, digested with LysCprotease and analysed by LC-MS/MS. The MH2+peptide that was fragmented to give the resulting MSspectra is shown on the right hand side of the spectrum (in green). The co-elution of two different peptides is shown by use of blue and red colours.

Discussion

Overall our results demonstrate that JMJD6 can undergo self-hydroxylation in recombinant form when produced in bacteria, and that self-hydroxylation is stimulated by 2OG addition. Evidence for hydroxylation at one site Lys167JMJD6 was also seen for JMJD6 isolated from human cells.

The amino acid analysis studies imply that the self-hydroxylation proceeds to give the product with the 5S-stereochemistry, i.e. the same as that observed for JMJD6 catalysed hydroxylation of COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
arginine serine
rich domain proteins (e.g.U2AF65).1,2 Thus, the JMJD6protein product differs from the 5R-hydroxylation stereochemistry observed for collagenlysylhydroxylation which is also catalysed by a 2OG dependent oxygenase.2,24,25 It should be noted that the combined amino acid analysis and proteomic studies imply that there are self-hydroxylation sites additional to Lys111JMJD6 and Lys167JMJD6. We thus suggest that wherever possible assignments of post-translational modification based on proteomic MS-based methods should be supplemented by amino acid or equivalent analyses. The amino acid analysis results also imply that the JMJD6 catalysed self-hydroxylation is active site mediated and is not a result of reactive oxidising species ‘leaking’ away from the COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
iron
centre. In this regard the JMJD6 catalysed self-hydroxylation is probably distinguished from that observed for other 2OG oxygenases to date, where the oxidised residues differ from the normal substrates. For example, in the case of FIH, a tryptophan residue is oxidised by self-oxidation,9 but a tryptophan residue has not been reported to be a substrate for FIH,26 (though this cannot be ruled out because FIH does catalyse hydroxylation of residues other than aspargine26,27). Thus it would seem that the JMJD6 catalysed self-hydroxylation is more likely to be biologically relevant than the self-oxidations so far observed for other 2OG oxygenases, in particular because the JMJD6 self-hydroxylation is observed in human cells.

Interestingly, peptides containing the hydroxylation sites Lys111JMJD6 and Lys167JMJD6 identified in JMJD6 recombinant protein were not observed to be substrates under our current assay conditions. Instead, two other peptides (corresponding to JMJD61–14 and JMJD6301–314) that were not observed to be hydroxylated in the JMJD6protein were observed to be substrates. Together these results reveal that factors other than the immediate sequence around the lysylhydroxylation site are important in determining selectivity as has been found for FIH,26 where relevant factors include secondary structure and overall protein stability.28 It should also be noted that for the self- hydroxylation of JMJD6, it is unclear if the reaction is intra- or inter-molecular (for either of the Lys111JMJD6, Lys167JMJD6 or other hydroxylation sites), and the available JMJD6 crystal structures29,30 do not rule out either possibility (Fig. 4).


Positions of self-hydroxylation sites within various 2OG oxygenases and ACCO. Ribbons representation of 2OG oxygenases and ACCO showing the distances in Angstorms (black dashes) between the iron centre (white sticks for protein derived ligands, green for 2OGcofactors and orange for iron or surrogate metal) and C-alpha atom of the self-hydroxylated residues (yellow sticks). a. JMJD6 (PDB ID3K2O), b. FIH (PDB ID1H2K), c. TauD (PDB ID1GQW), d. ACCO (PDB ID1W9Y), e. ABH3 (PDB ID2IUW), f. AlkB (PDB ID2FD8). In the case of ACCOhydroxylation was not observed directly but is proposed to lead to backbone fragmentation. Note that the distances to the two residues as being hydroxylated in JMJD6 are substantially longer than in the other structures and other JMJD6 residues (than Lys111JMJD6 or Lys167JMJD6) are likely hydroxylated (see text).
Fig. 4 Positions of self-hydroxylation sites within various 2OG oxygenases and ACCO. Ribbons representation of 2OG oxygenases and ACCO showing the distances in Angstorms (black dashes) between the COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
iron
centre (white sticks for protein derived ligands, green for 2OGcofactors and orange for COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
iron
or surrogate metal) and C-alpha atom of the self-hydroxylated residues (yellow sticks). a. JMJD6 (PDB ID3K2O), b. FIH (PDB ID1H2K), c. TauD (PDB ID1GQW), d. ACCO (PDB ID1W9Y), e. ABH3 (PDB ID2IUW), f. AlkB (PDB ID2FD8). In the case of ACCOhydroxylation was not observed directly but is proposed to lead to backbone fragmentation. Note that the distances to the two residues as being hydroxylated in JMJD6 are substantially longer than in the other structures and other JMJD6 residues (than Lys111JMJD6 or Lys167JMJD6) are likely hydroxylated (see text).

For other 2OG oxygenases9–11,14 and the related enzymeACC oxidase,15 the identified oxidative modifications have all been observed to occur within ∼10 Å of the active site COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
iron
(Fig. 4), implying that they are formed directly by reaction of active site COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
iron
bound oxidising species or leakage of reactive oxidising species (e.g.H2O2, O2˙) away from the active site. However, in the case of JMJD6 the identified sites of hydroxylation at Lys111JMJD6, Lys167JMJD6 are both ∼22 Å away from the active site COMPOUND LINKS

Read more about this on ChemSpider

Download mol file of compound
iron
as shown in Fig. 4. Lys111JMJD6 and Lys167JMJD6 are conserved in JMJD6 across higher eukaryotes (Fig. S6), with deviations in residues observed in the corresponding Lys167JMJD6 position in lower organisms. Both Lys111JMJD6 and Lys167JMJD6 are located in loop regions that on the basis of the crystallographic analysis are likely to be relatively flexible which may explain how these residues are able to fit into the active site to be hydroxylated.

We also observed that self-hydroxylation and uncoupled turnover occur in the presence of a JMJD6 substrate, U2AF65. With the available techniques, it is difficult to study the effects of, at least, several self-hydroxylation events on substrate hydroxylation; however, self-hydroxylation may reduce the extent of substrate hydroxylation by simple competition. It is now of interest to see if the extent of self-hydroxylations varies with different JMJD6 substrates and if self-hydroxylation has any biological role. As suggested for other 2OG oxygenases, it is possible that self-hydroxylation of JMJD6 is involved in (likely negative) feedback mechanisms, though at this stage there is no evidence for the in vivo relevance of self hydroxylation.9JMJD6 has been shown to play a role in the regulation of splicing of genes including vascular endothelial growth factor (VEGF) receptor 1,3 but the exact role of JMJD6 catalysed hydroxylation is unclear. At present the available evidence suggests that, as proposed for hydroxylation catalysed by FIH,26–28,31,32JMJD6 may have multiple hydroxylation substrates and that the sum of many hydroxylation events may be regulatory in a manner that is context dependent. It is of interest that hydroxylation mediated modulation of FIH and ankyrin repeat domain substrate interactions is proposed to regulate the role of FIH in hydroxylation of C-terminal transcriptional activity domain of HIF.31–33 The occurrence of self-hydroxylation by JMJD6 introduces another layer of complexity, potentially applying to other 2OG oxygenases.

Thus, it is now of interest to study self-hydroxylation by other human 2OG oxygenases including the collagenlysyl hydroxylase, and the JmjC containing histoneNε-methyl lysine demethylases. Nε-Methyl lysine formation is a common post-translational modification as observed by proteomic studies,34 thus self-demethylation is also a possibility for some 2OG oxygenases. If indeed self-oxidation plays a regulatory role, its modulation should be taken into consideration when considering the phenotypic effects of small molecule inhibitors targeting 2OG oxygenases subject to self-oxidation.

Acknowledgements

We thank Holger Kramer, Benedikt Kessler and Matthias Mann for helpful discussions on interpretation of mass spectrometric results. (Michael Nielsen and Matthias Mann performed the mass spectrometry reported in this work.) We thank Timothy D. W. Claridge for discussion on NMR work. This work was funded by DFG grant BO1748–3 awarded to A. B. and the Biotechnology and Biological Sciences Research Council, the European Union and the Wellcome Trust. C. W. was funded by a Glasstone Fellowship.

References

  1. C. J. Webby, A. Wolf, N. Gromak, M. Dreger, H. Kramer, B. Kessler, M. L. Nielsen, C. Schmitz, D. S. Butler, J. R. Yates, III, C. M. Delahunty, P. Hahn, A. Lengeling, M. Mann, N. J. Proudfoot, C. J. Schofield and A. Bottger, Science, 2009, 325, 90–93 CrossRef CAS.
  2. M. Mantri, N. D. Loik, R. B. Hamed, T. D. W. Claridge, J. S. O. McCullagh and C. J. Schofield, ChemBioChem, 2011, 12, 531–534 CrossRef CAS.
  3. J.-N. Boeckel, V. Guarani, M. Koyanagi, T. Roexe, A. Lengeling, R. T. Schermuly, P. Gellert, T. Braun, A. Zeiher and S. Dimmeler, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, 3276–3281 CrossRef CAS.
  4. B. Chang, Y. Chen, Y. Zhao and R. K. Bruick, Science, 2007, 318, 444–447 CrossRef CAS.
  5. C. Loenarz and C. J. Schofield, Nat. Chem. Biol., 2008, 4, 152–156 CrossRef CAS.
  6. A. Wolf, C. Schmitz and A. Böettger, EMBO Rep., 2007, 8, 465–469 CrossRef CAS.
  7. N. R. Rose, M. A. McDonough, O. N. F. King, A. Kawamura and C. J. Schofield, Chem. Soc. Rev., 2011, 40, 4364–4397 RSC.
  8. D. F. Counts, G. J. Cardinale and S. Udenfriend, Proc. Natl. Acad. Sci. U. S. A., 1978, 75, 2145–2149 CrossRef CAS.
  9. Y. H. Chen, L. M. Comeaux, S. J. Eyles and M. J. Knapp, Chem. Commun., 2008, 4768–4770 RSC.
  10. M. J. Ryle, A. Liu, R. B. Muthukumaran, R. Y. Ho, K. D. Koehntop, J. McCracken, L. Que, Jr. and R. P. Hausinger, Biochemistry, 2003, 42, 1854–1862 CrossRef CAS.
  11. A. Liu, R. Y. N. Ho, L. Que, M. J. Ryle, B. S. Phinney and R. P. Hausinger, J. Am. Chem. Soc., 2001, 123, 5126–5127 CrossRef CAS.
  12. R. W. D. Welford, I. Schlemminger, L. A. McNeill, K. S. Hewitson and C. J. Schofield, J. Biol. Chem., 2003, 278, 10157–10161 CrossRef CAS.
  13. M. Cikala, O. Alexandrova, C. David, M. Proschel, B. Stiening, P. Cramer and A. Böttger, BMC Cell Biol., 2004, 5, 26–36 CrossRef.
  14. O. Sundheim, C. B. Vagbo, M. Bjoras, M. M. L. Sousa, V. Talstad, P. A. Aas, F. Drablos, H. E. Krokan, J. A. Tainer and G. Slupphaug, EMBO J., 2006, 25, 3389–3397 CrossRef CAS.
  15. Z. Zhang, J.-S. Ren, I. J. Clifton and C. J. Schofield, Chem. Biol., 2004, 11, 1383–1394 CrossRef CAS.
  16. H. Yamaguchi and W. A. Hendrickson, Nature, 1996, 384, 484–489 CrossRef CAS.
  17. S. Fang, J. P. Jensen, R. L. Ludwig, K. H. Vousden and A. M. Weissman, J. Biol. Chem., 2000, 275, 8945–8951 CrossRef CAS.
  18. J. Tang, L. K. Qu, J. Zhang, W. Wang, J. S. Michaelson, Y. Y. Degenhardt, W. S. El-Deiry and X. Yang, Nat. Cell Biol., 2006, 8, 855–862 CrossRef CAS.
  19. P. J. Sabourin and L. L. Bieber, J. Biol. Chem., 1982, 257, 7460–7467 CAS.
  20. R. Myllylä, K. Majamaa, V. Günzler, H. M. Hanauske-Abel and K. I. Kivirikko, J. Biol. Chem., 1984, 259, 5403–5405 Search PubMed.
  21. E. Flashman, S. L. Davies, K. K. Yeoh and C. J. Schofield, Biochem. J., 2010, 427, 135–142 CrossRef CAS.
  22. K. J. Nytko, N. Maeda, P. Schlāfli, P. Spielmann, R. H. Wenger and D. P. Stiehl, Blood, 2011 Search PubMed.
  23. J. Cox and M. Mann, Nat. Biotechnol., 2008, 26, 1367–1372 CrossRef CAS.
  24. B. Witkop, Experientia, 1956, 12, 372–374 CrossRef CAS.
  25. B. Witkop, Chemical Society (London), 1955, 3, 60–82 Search PubMed.
  26. M. Yang, W. Ge, R. Chowdhury, T. D. W. Claridge, H. B. Kramer, B. Schmierer, M. A. McDonough, L. Gong, B. M. Kessler, P. J. Ratcliffe, M. L. Coleman and C. J. Schofield, J. Biol. Chem., 2011, 286, 7648–7660 CrossRef CAS.
  27. M. Yang, R. Chowdhury, W. Ge, R. B. Hamed, M. A. McDonough, T. D. W. Claridge, B. M. Kessler, M. E. Cockman, P. J. Ratcliffe and C. J. Schofield, FEBS J., 2011, 278, 1086–1097 CrossRef CAS.
  28. L. Kelly, M. A. McDonough, M. L. Coleman, P. J. Ratcliffe and C. J. Schofield, Mol. BioSyst., 2009, 5, 52–58 RSC.
  29. M. Mantri, T. Krojer, E. A. Bagg, C. J. Webby, D. S. Butler, G. Kochan, K. L. Kavanagh, U. Oppermann, M. A. McDonough and C. J. Schofield, J. Mol. Biol., 2010, 401, 211–222 CrossRef CAS.
  30. X. Hong, J. Zang, J. White, C. Wang, C.-H. Pan, R. Zhao, R. C. Murphy, S. Dai, P. Henson, J. W. Kappler, J. Hagman and G. Zhang, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 14568–14572 CrossRef CAS.
  31. M. L. Coleman, M. A. McDonough, K. S. Hewitson, C. Coles, J. Mecinović, M. Edelmann, K. M. Cook, M. E. Cockman, D. E. Lancaster, B. M. Kessler, N. J. Oldham, P. J. Ratcliffe and C. J. Schofield, J. Biol. Chem., 2007, 282, 24027–24038 CrossRef CAS.
  32. M. E. Cockman, J. D. Webb and P. J. Ratcliffe, Ann. N. Y. Acad. Sci., 2009, 1177, 9–18 CrossRef CAS.
  33. B. Schmierer, B. Novak and C. Schofield, BMC Syst. Biol., 2010, 4, 139–156 CrossRef.
  34. M. Vermeulen, H. C. Eberl, F. Matarese, H. Marks, S. Denissov, F. Butter, K. K. Lee, J. V. Olsen, A. A. Hyman, H. G. Stunnenberg and M. Mann, Cell, 2010, 142, 967–980 CrossRef CAS.

Footnotes

This article is part of a MedChemComm web themed issue on epigenetics.
Electronic supplementary information (ESI) available. See DOI: 10.1039/c1md00225b

This journal is © The Royal Society of Chemistry 2012