Guardians of the Genome: Iron–Sulfur Proteins in the Nucleus
"> Figure 1
<p>A comparison of solvent-exposed and buried Fe-S clusters in proteins. The structures (left) and close-ups of the Fe-S binding domains (square, right). (<b>A</b>) X-ray structure of human xanthine oxidase (XDH) showing the two [2Fe-2S] clusters deeply buried within the protein structure (PDB 2CKJ (<a href="https://www.rcsb.org/structure/2CKJ" target="_blank">https://www.rcsb.org/structure/2CKJ</a>, accessed on 30 September 2024)). (<b>B</b>) Cryo-EM structure of the human ribosome recycling factor ABCE1 showing the two [4Fe-4S] clusters situated in solvent-exposed loops at the N-terminal domain (PDB 7A09 (<a href="https://www.rcsb.org/structure/7A09" target="_blank">https://www.rcsb.org/structure/7A09</a>, accessed on 1 October 2024)). The clusters in ABCE1 are prone to degradation [<a href="#B102-inorganics-12-00316" class="html-bibr">102</a>]. Each cluster in ABCE1 is coordinated by three nearby Cys residues and one distant Cys residue in the primary sequence, suggesting a structural role for these clusters [<a href="#B103-inorganics-12-00316" class="html-bibr">103</a>]. (<b>C</b>) Cryo-EM structure of human DNA primase subunits, PRIM1 (green) and PRIM2 (blue) (PDB 8QJ7 (<a href="https://www.rcsb.org/structure/8QJ7" target="_blank">https://www.rcsb.org/structure/8QJ7</a>, accessed on 2 October 2024)). The [4Fe-4S] cluster bound to PRIM2 locates at the junction of the N-terminal and C-terminal domains close to the protein surface. PRIM1 is proposed to serve as an adaptor for the recognition of PRIM2 by the CTC to facilitate its metallation. (<b>D</b>) Cryo-EM structure of the yeast elongator complex proteins ELP3 (green) and ELP4 (yellow) (PDB 8ASV (<a href="https://www.rcsb.org/structure/8ASV" target="_blank">https://www.rcsb.org/structure/8ASV</a>, accessed on 2 October 2024)). The radical SAM enzyme ELP3 has a [4Fe-4S](Cys)<sub>3</sub> cluster, which is exposed to solvent via two channels [<a href="#B103-inorganics-12-00316" class="html-bibr">103</a>]. ELP4 is proposed to serve as an adaptor for the recognition of ELP3 by the CTC to facilitate its metallation.</p> "> Figure 2
<p>The essential roles of nuclear Fe-S cluster-binding proteins in DNA replication, telomere maintenance, and mitosis. (<b>A</b>) A representation of a eukaryotic DNA replication fork. The mini-chromosome maintenance (MCM2-7) helicase complex unwinds DNA at replication origins, separating the double helix to generate the leading and lagging strands. All replicative DNA polymerases, Pol α, δ, and ε, bind a [4Fe-4S] cluster. DNA replication initiates with an RNA primer (orange) synthesized by DNA primase, with the regulatory subunit PRIM2 also binding a [4Fe-4S] cluster. Pol ε further extends the primer on the leading strand, while Pol δ does so on the lagging strand. The proliferating cell nuclear antigen sliding clamp PCNA enhances the processivity of DNA polymerases and facilitates DNA repair. The [4Fe-4S] cluster DNA helicases FANCJ and DDX11 unwind G4s to prevent stalling of the replisome. DNA lesions obstruct replication, triggering monoubiquitylation of PCNA, which recruits the DNA damage bypass DNA polymerase Pol ζ. The catalytic subunit of Pol ζ also binds a [4Fe-4S] cluster. Finally, the [4Fe-4S] cluster nuclease–helicase DNA2 cleaves exposed single-stranded DNA ends from Okazaki fragments and stalled replication forks. (<b>B</b>) Under Fe-sufficient conditions, the MCM helicase moves in opposite directions from the activated origin of replication. This results in the formation of two functional replication forks, enabling DNA synthesis. Under iron-depleted conditions, apo NCOA4 accumulates in the nucleus and binds to the helicase subunit MCM7, hindering MCM helicase activity and inactivating replication origins. This ensures that DNA synthesis only occurs when there is a sufficient pool of metallated Fe-S enzymes, thereby maintaining genomic integrity and preventing replication under suboptimal conditions. (<b>C</b>) The [4Fe-4S] cluster helicase RTEL1 unwinds G4s and R-loops at the telomeric repeats, facilitating telomere replication by telomerase and preventing telomere shortening. (<b>D</b>) During the late stages of mitosis (metaphase, anaphase, and telophase), the nuclear [4Fe-4S] cluster mitotic factor KIF4A binds the arms of condensed chromosomes. As a kinesin, KIF4A moves along microtubules to mobilize cargoes. KIF4A localizes at the central spindle, accumulating at the spindle midzone and midbody. The close-up image on the bottom illustrates the colocalization of the CIA targeting factors MMS19 and CIAO2B with KIF4A during mitosis, suggesting in situ metallation of KIF4A during function.</p> "> Figure 3
<p>The Fe-S cluster proteins within the six major DNA repair pathways: base excision repair (BER), nucleotide excision repair (NER), mismatch repair (MMR), homologous recombination (HR), non-homologous end joining (NHEJ), and translesion DNA synthesis (TLS). A schematic representation of DNA lesions and the corresponding repair pathways. The key proteins involved in each pathway are organized according Knijnenburg et al. [<a href="#B172-inorganics-12-00316" class="html-bibr">172</a>]. DNA repair proteins that bind [4Fe-4S] clusters are shown in a blue square at the bottom. * While DDX11 and EXO5 are not formally categorized under HR, their functions align with this repair mechanism.</p> ">
Abstract
:1. Introduction
Participates In | Name | Consolidated Localization | Fe-S Cluster Type |
---|---|---|---|
DNA replication/DNA repair | PRIM2 | Nucleus | [4Fe-4S] (Cys)4 [8,9] |
POLA1 | Nucleus, Cytosol [10] | [4Fe-4S] (Cys)4 [11] | |
POLD1 | Nucleus, Cytosol [12] | [4Fe-4S] (Cys)4 [11] | |
POLE | Nucleus | [4Fe-4S] (Cys)4 [11,13] | |
REV3L | Nucleus, Cytosol [14] | [4Fe-4S] (Cys)4 [11] | |
FANCJ | Nucleus, Cytosol [15,16] | [4Fe-4S] (Cys)4 [4,17] | |
DDX11 | Nucleus [18] | [4Fe-4S] (Cys)4 [19] | |
RTEL1 | Nucleus [16] | [4Fe-4S] (Cys)4 [20] | |
XPD | Nucleus, Cytosol [16,21] | [4Fe-4S] (Cys)4 [4] | |
DNA2 | Nucleus, Mitochondria [22] | [4Fe-4S] (Cys)4 [23,24] | |
EXO5 | Nucleus, Cytosol [25] | [4Fe-4S] (Cys)4 [25] | |
MUTYH | Nucleus, Mitochondria [26] | [4Fe-4S] (Cys)4 [27,28] | |
NTHL1 | Nucleus, Mitochondria [29] | [4Fe-4S] (Cys)4 [30,31] | |
RNA transactions | ELP3 | Nucleus, Cytosol [32] | [4Fe-4S] (Cys)3 SAM [33,34] |
RPC6 | Nucleus [35] | [4Fe-4S] (Cys)4 [36,37] | |
CPSF4 | Nucleus [38] | [2Fe-2S] (Cys)3 (His)? [39,40] | |
TYW1 | Nucleus, Cytosol | [4Fe-4S] (Cys)3 SAM [34,41] | |
TYW1B | Nucleus | [4Fe-4S] (Cys)3 SAM [34,41] | |
CDK5RAP1 | Nucleus, Mitochondria, Cytosol [42] | [4Fe-4S] (Cys)4, [4Fe-4S] (Cys)3 SAM [34,43] | |
Mitosis | KIF4A | Nucleus [44] | [4Fe-4S] (Cys)4? [45] |
KIF4B | Nucleus [44] | [4Fe-4S] (Cys)4? [45] | |
Iron metabolism | NCOA4 | Nucleus, Cytosol [46] | [3Fe-4S] (Cys)4? [47,48] |
FBXL5 | Nucleus, Perinuclear region [49] | [2Fe-2S] (Cys)4 [50] | |
Post-translational modifications | DPH1 | Nucleus, Cytosol [51] | [4Fe-4S] (Cys)3 SAM [52] |
DPH2 | Nucleus, Cytosol | [4Fe-4S] (Cys)3 SAM [52] | |
ATE1 | Nucleus, Cytosol [53] | [4Fe-4S] (Cys)4 [54] | |
Respiration | SDHB | Mitochondria, Nucleus [55] | [2Fe-2S] (Cys)4, [3Fe-4S] (Cys)3, [4Fe-4S] (Cys)4 [55] |
Unknown | RFESD | Nucleus | [2Fe-2S] (Cys)2 (His)2 |
CIA | CIAPIN1 | Nucleus, Mitochondria, Cytosol [56] | [2Fe-2S] (Cys)4, [4Fe-4S] (Cys)4? [57,58] |
BOLA2 | Nucleus, Cytosol [59] | [2Fe-2S] * Ligands shared with GLRX3 [60,61] | |
GLRX3 | Nucleus, Cytosol [62] | [2Fe-2S] * Ligands shared with BOLA2 [60,61] | |
NUBP2 | Nucleus, Cytosol [63] | [4Fe-4S] * Ligands shared with NUBP1 [64,65] | |
ISC | NFU1 | Nucleus, Mitochondria, Cytosol [66] | [4Fe-4S] * Ligands shared between dimers [66,67] |
NFS1 | Nucleus, Mitochondria, Cytosol [68] | [2Fe-2S] * Ligands shared with ISCU [69,70] | |
GLRX2 | Nucleus, Mitochondria [71] | [2Fe-2S] * Ligands shared between dimers [72] |
2. Metallation of Nuclear Fe-S Cluster Proteins
3. Fe-S Cluster Proteins Involved in Nuclear DNA Transactions
3.1. Fe-S Clusters and DNA Replication
3.1.1. DNA Polymerases
3.1.2. DNA Primase
3.1.3. Fe-S Clusters and DNA Replication Origins
3.1.4. Fe-S Clusters in Helicase Activity
3.1.5. Fe-S Clusters and Telomere Maintenance
3.2. Fe-S Clusters and DNA Repair
3.3. Fe-S Clusters, Transcription, and Nuclear RNA Transactions
3.3.1. Fe-S Cluster-Binding Fe-Dependent Transcriptional Regulators
3.3.2. RNA Polymerase III
3.3.3. Transcription Factor IIH
3.3.4. RNA-Modifying Proteins
4. Fe-S Cluster Proteins in Mitosis
5. Other Fe-S Cluster Proteins in the Nucleus
6. Conclusions and Open Questions
Supplementary Materials
Author Contributions
Funding
Data Availability Statement
Acknowledgments
Conflicts of Interest
References
- Lill, R.; Broderick, J.B.; Dean, D.R. Special issue on iron-sulfur proteins: Structure, function, biogenesis and diseases. Biochim. Biophys. Acta 2015, 1853, 1251–1252. [Google Scholar] [CrossRef] [PubMed]
- Braymer, J.J.; Freibert, S.A.; Rakwalska-Bange, M.; Lill, R. Mechanistic concepts of iron-sulfur protein biogenesis in Biology. Biochim. Et. Biophys. Acta (BBA)-Mol. Cell Res. 2021, 1868, 118863. [Google Scholar] [CrossRef] [PubMed]
- Bak, D.W.; Weerapana, E. Proteomic strategies to interrogate the Fe-S proteome. Biochim. Biophys. Acta Mol. Cell Res. 2024, 1871, 119791. [Google Scholar] [CrossRef] [PubMed]
- Rudolf, J.; Makrantoni, V.; Ingledew, W.J.; Stark, M.J.; White, M.F. The DNA repair helicases XPD and FancJ have essential iron-sulfur domains. Mol. Cell 2006, 23, 801–808. [Google Scholar] [CrossRef]
- Thul, P.J.; Åkesson, L.; Wiking, M.; Mahdessian, D.; Geladaki, A.; Ait Blal, H.; Alm, T.; Asplund, A.; Björk, L.; Breckels, L.M. A subcellular map of the human proteome. Science 2017, 356, eaal3321. [Google Scholar] [CrossRef]
- Consortium, T.U. UniProt: The universal protein knowledgebase in 2023. Nucleic Acids Res. 2023, 51, D523–D531. [Google Scholar] [CrossRef]
- Binder, J.X.; Pletscher-Frankild, S.; Tsafou, K.; Stolte, C.; O’Donoghue, S.I.; Schneider, R.; Jensen, L.J. COMPARTMENTS: Unification and visualization of protein subcellular localization evidence. Database 2014, 2014, bau012. [Google Scholar] [CrossRef]
- Klinge, S.; Hirst, J.; Maman, J.D.; Krude, T.; Pellegrini, L. An iron-sulfur domain of the eukaryotic primase is essential for RNA primer synthesis. Nat. Struct. Mol. Biol. 2007, 14, 875–877. [Google Scholar] [CrossRef]
- Weiner, B.E.; Huang, H.; Dattilo, B.M.; Nilges, M.J.; Fanning, E.; Chazin, W.J. An iron-sulfur cluster in the C-terminal domain of the p58 subunit of human DNA primase. J. Biol. Chem. 2007, 282, 33444–33451. [Google Scholar] [CrossRef]
- Starokadomskyy, P.; Gemelli, T.; Rios, J.J.; Xing, C.; Wang, R.C.; Li, H.; Pokatayev, V.; Dozmorov, I.; Khan, S.; Miyata, N. DNA polymerase-α regulates the activation of type I interferons through cytosolic RNA: DNA synthesis. Nat. Immunol. 2016, 17, 495–504. [Google Scholar] [CrossRef]
- Netz, D.J.; Stith, C.M.; Stümpfig, M.; Köpf, G.; Vogel, D.; Genau, H.M.; Stodola, J.L.; Lill, R.; Burgers, P.M.; Pierik, A.J. Eukaryotic DNA polymerases require an iron-sulfur cluster for the formation of active complexes. Nat. Chem. Biol. 2012, 8, 125–132. [Google Scholar] [CrossRef] [PubMed]
- Chea, J.; Zhang, S.; Zhao, H.; Zhang, Z.; Lee, E.Y.; Darzynkiewicz, Z.; Lee, M.Y. Spatiotemporal recruitment of human DNA polymerase delta to sites of UV damage. Cell Cycle 2012, 11, 2885–2895. [Google Scholar] [CrossRef] [PubMed]
- Roske, J.J.; Yeeles, J.T. Structural basis for processive daughter-strand synthesis and proofreading by the human leading-strand DNA polymerase Pol ε. Nat. Struct. Mol. Biol. 2024, 1–11. [Google Scholar] [CrossRef]
- Ræder, S.B.; Nepal, A.; Bjørås, K.Ø.; Seelinger, M.; Kolve, R.S.; Nedal, A.; Müller, R.; Otterlei, M. APIM-mediated REV3L–PCNA interaction important for error free TLS Over UV-induced DNA lesions in human cells. Int. J. Mol. Sci. 2018, 20, 100. [Google Scholar] [CrossRef]
- Cantor, S.B.; Bell, D.W.; Ganesan, S.; Kass, E.M.; Drapkin, R.; Grossman, S.; Wahrer, D.C.; Sgroi, D.C.; Lane, W.S.; Haber, D.A. BACH1, a novel helicase-like protein, interacts directly with BRCA1 and contributes to its DNA repair function. Cell 2001, 105, 149–160. [Google Scholar] [CrossRef]
- Seki, M.; Takeda, Y.; Iwai, K.; Tanaka, K. IOP1 protein is an external component of the human cytosolic iron-sulfur cluster assembly (CIA) machinery and functions in the MMS19 protein-dependent CIA pathway. J. Biol. Chem. 2013, 288, 16680–16689. [Google Scholar] [CrossRef]
- Wu, Y.; Sommers, J.A.; Suhasini, A.N.; Leonard, T.; Deakyne, J.S.; Mazin, A.V.; Shin-Ya, K.; Kitao, H.; Brosz, R.M., Jr. Fanconi anemia group J mutation abolishes its DNA repair function by uncoupling DNA translocation from helicase activity or disruption of protein-DNA complexes. Blood J. Am. Soc. Hematol. 2010, 116, 3780–3791. [Google Scholar] [CrossRef]
- Parish, J.L.; Rosa, J.; Wang, X.; Lahti, J.M.; Doxsey, S.J.; Androphy, E.J. The DNA helicase ChlR1 is required for sister chromatid cohesion in mammalian cells. J. Cell Sci. 2006, 119, 4857–4865. [Google Scholar] [CrossRef]
- Simon, A.K.; Kummer, S.; Wild, S.; Lezaja, A.; Teloni, F.; Jozwiakowski, S.K.; Altmeyer, M.; Gari, K. The iron–sulfur helicase DDX11 promotes the generation of single-stranded DNA for CHK1 activation. Life Sci. Alliance 2020, 3. [Google Scholar] [CrossRef]
- Landry, A.P.; Ding, H. The N-Terminal Domain of Human DNA Helicase Rtel1 Contains a Redox Active Iron-Sulfur Cluster. BioMed Res. Int. 2014, 2014, 285791. [Google Scholar] [CrossRef]
- Ito, S.; Tan, L.J.; Andoh, D.; Narita, T.; Seki, M.; Hirano, Y.; Narita, K.; Kuraoka, I.; Hiraoka, Y.; Tanaka, K. MMXD, a TFIIH-independent XPD-MMS19 protein complex involved in chromosome segregation. Mol. Cell 2010, 39, 632–640. [Google Scholar] [CrossRef] [PubMed]
- Duxin, J.P.; Dao, B.; Martinsson, P.; Rajala, N.; Guittat, L.; Campbell, J.L.; Spelbrink, J.N.; Stewart, S.A. Human Dna2 is a nuclear and mitochondrial DNA maintenance protein. Mol. Cell. Biol. 2009, 29, 4274–4282. [Google Scholar] [CrossRef] [PubMed]
- Pokharel, S.; Campbell, J.L. Cross talk between the nuclease and helicase activities of Dna2: Role of an essential iron–sulfur cluster domain. Nucleic Acids Res. 2012, 40, 7821–7830. [Google Scholar] [CrossRef]
- Mariotti, L.; Wild, S.; Brunoldi, G.; Piceni, A.; Ceppi, I.; Kummer, S.; Lutz, R.E.; Cejka, P.; Gari, K. The iron–sulphur cluster in human DNA2 is required for all biochemical activities of DNA2. Commun. Biol. 2020, 3, 322. [Google Scholar] [CrossRef]
- Sparks, J.L.; Kumar, R.; Singh, M.; Wold, M.S.; Pandita, T.K.; Burgers, P.M. Human exonuclease 5 is a novel sliding exonuclease required for genome stability. J. Biol. Chem. 2012, 287, 42773–42783. [Google Scholar] [CrossRef]
- Komine, K.; Shimodaira, H.; Takao, M.; Soeda, H.; Zhang, X.; Takahashi, M.; Ishioka, C. Functional complementation assay for 47 MUTYH variants in a MutY-disrupted Escherichia coli strain. Hum. Mutat. 2015, 36, 704–711. [Google Scholar] [CrossRef]
- Chepanoske, C.L.; Golinelli, M.-P.; Williams, S.D.; David, S.S. Positively Charged Residues within the Iron–Sulfur Cluster Loop of E. coli MutY Participate in Damage Recognition and Removal. Arch. Biochem. Biophys. 2000, 380, 11–19. [Google Scholar] [CrossRef]
- Nuñez, N.N.; Majumdar, C.; Lay, K.T.; David, S.S. Fe–S clusters and MutY base excision repair glycosylases: Purification, kinetics, and DNA affinity measurements. In Methods in Enzymology; Elsevier: Amsterdam, The Netherlands, 2018; Volume 599, pp. 21–68. [Google Scholar]
- Ikeda, S.; Kohmoto, T.; Tabata, R.; Seki, Y. Differential intracellular localization of the human and mouse endonuclease III homologs and analysis of the sorting signals. DNA Repair. 2002, 1, 847–854. [Google Scholar] [CrossRef]
- Kuo, C.-F.; McRee, D.E.; Fisher, C.L.; O’Handley, S.F.; Cunningham, R.P.; Tainer, J.A. Atomic structure of the DNA repair [4Fe-4S] enzyme endonuclease III. Science 1992, 258, 434–440. [Google Scholar] [CrossRef]
- Carroll, B.L.; Zahn, K.E.; Hanley, J.P.; Wallace, S.S.; Dragon, J.A.; Doublié, S. Caught in motion: Human NTHL1 undergoes interdomain rearrangement necessary for catalysis. Nucleic Acids Res. 2021, 49, 13165–13178. [Google Scholar] [CrossRef]
- Kim, J.-H.; Lane, W.S.; Reinberg, D. Human Elongator facilitates RNA polymerase II transcription through chromatin. Proc. Natl. Acad. Sci. USA 2002, 99, 1241–1246. [Google Scholar] [CrossRef] [PubMed]
- Paraskevopoulou, C.; Fairhurst, S.A.; Lowe, D.J.; Brick, P.; Onesti, S. The elongator subunit Elp3 contains a Fe4S4 cluster and binds S-adenosylmethionine. Mol. Microbiol. 2006, 59, 795–806. [Google Scholar] [CrossRef]
- Kimura, S.; Suzuki, T. Iron-sulfur proteins responsible for RNA modifications. Biochim. Biophys. Acta 2015, 1853, 1272–1283. [Google Scholar] [CrossRef] [PubMed]
- Ramsay, E.P.; Abascal-Palacios, G.; Daiß, J.L.; King, H.; Gouge, J.; Pilsl, M.; Beuron, F.; Morris, E.; Gunkel, P.; Engel, C. Structure of human RNA polymerase III. Nat. Commun. 2020, 11, 6409. [Google Scholar] [CrossRef] [PubMed]
- Blombach, F.; Salvadori, E.; Fouqueau, T.; Yan, J.; Reimann, J.; Sheppard, C.; Smollett, K.L.; Albers, S.V.; Kay, C.W.; Thalassinos, K. Archaeal TFEα/β is a hybrid of TFIIE and the RNA polymerase III subcomplex hRPC62/39. Elife 2015, 4, e08378. [Google Scholar] [CrossRef]
- Li, L.; Yu, Z.; Zhao, D.; Ren, Y.; Hou, H.; Xu, Y. Structure of human RNA polymerase III elongation complex. Cell Res. 2021, 31, 791–800. [Google Scholar] [CrossRef]
- Chen, W.; Qin, L.; Wang, S.; Li, M.; Shi, D.; Tian, Y.; Wang, J.; Fu, L.; Li, Z.; Guo, W. CPSF4 activates telomerase reverse transcriptase and predicts poor prognosis in human lung adenocarcinomas. Mol. Oncol. 2014, 8, 704–716. [Google Scholar] [CrossRef]
- Shimberg, G.D.; Michalek, J.L.; Oluyadi, A.A.; Rodrigues, A.V.; Zucconi, B.E.; Neu, H.M.; Ghosh, S.; Sureschandra, K.; Wilson, G.M.; Stemmler, T.L.; et al. Cleavage and polyadenylation specificity factor 30: An RNA-binding zinc-finger protein with an unexpected 2Fe-2S cluster. Proc. Natl. Acad. Sci. USA 2016, 113, 4700–4705. [Google Scholar] [CrossRef]
- Pritts, J.D.; Hursey, M.S.; Michalek, J.L.; Batelu, S.; Stemmler, T.L.; Michel, S.L. Unraveling the RNA Binding Properties of the Iron–Sulfur Zinc Finger Protein CPSF30. Biochemistry 2020, 59, 970–982. [Google Scholar] [CrossRef]
- Young, A.P.; Bandarian, V. TYW1: A radical SAM enzyme involved in the biosynthesis of wybutosine bases. In Methods in Enzymology; Elsevier: Amsterdam, The Netherlands, 2018; Volume 606, pp. 119–153. [Google Scholar]
- Reiter, V.; Matschkal, D.M.; Wagner, M.; Globisch, D.; Kneuttinger, A.C.; Muller, M.; Carell, T. The CDK5 repressor CDK5RAP1 is a methylthiotransferase acting on nuclear and mitochondrial RNA. Nucleic Acids Res. 2012, 40, 6235–6240. [Google Scholar] [CrossRef]
- Pierrel, F.; Douki, T.; Fontecave, M.; Atta, M. MiaB protein is a bifunctional radical-S-adenosylmethionine enzyme involved in thiolation and methylation of tRNA. J. Biol. Chem. 2004, 279, 47555–47563. [Google Scholar] [CrossRef] [PubMed]
- Lee, Y.M.; Lee, S.; Lee, E.; Shin, H.; Hahn, H.; Choi, W.; Kim, W. Human kinesin superfamily member 4 is dominantly localized in the nuclear matrix and is associated with chromosomes during mitosis. Biochem. J. 2001, 360, 549–556. [Google Scholar] [CrossRef] [PubMed]
- Ben-Shimon, L.; Paul, V.D.; David-Kadoch, G.; Volpe, M.; Stümpfig, M.; Bill, E.; Mühlenhoff, U.; Lill, R.; Ben-Aroya, S. Fe-S cluster coordination of the chromokinesin KIF4A alters its subcellular localization during mitosis. J. Cell Sci. 2018, 131, jcs211433. [Google Scholar] [CrossRef]
- Bellelli, R.; Castellone, M.D.; Guida, T.; Limongello, R.; Dathan, N.A.; Merolla, F.; Cirafici, A.M.; Affuso, A.; Masai, H.; Costanzo, V. NCOA4 transcriptional coactivator inhibits activation of DNA replication origins. Mol. Cell 2014, 55, 123–137. [Google Scholar] [CrossRef]
- Kuno, S.; Iwai, K. Oxygen modulates iron homeostasis by switching iron sensing of NCOA4. J. Biol. Chem. 2023, 299. [Google Scholar] [CrossRef]
- Zhao, H.; Lu, Y.; Zhang, J.; Sun, Z.; Cheng, C.; Liu, Y.; Wu, L.; Zhang, M.; He, W.; Hao, S. NCOA4 requires a [3Fe-4S] to sense and maintain the iron homeostasis. J. Biol. Chem. 2024, 300, 105612. [Google Scholar] [CrossRef]
- Vinas-Castells, R.; Frias, A.; Robles-Lanuza, E.; Zhang, K.; Longmore, G.D.; Garcia de Herreros, A.; Diaz, V.M. Nuclear ubiquitination by FBXL5 modulates Snail1 DNA binding and stability. Nucleic Acids Res. 2013, 42, 1079–1094. [Google Scholar] [CrossRef]
- Wang, H.; Shi, H.; Rajan, M.; Canarie, E.R.; Hong, S.; Simoneschi, D.; Pagano, M.; Bush, M.F.; Stoll, S.; Leibold, E.A. FBXL5 regulates IRP2 stability in iron homeostasis via an oxygen-responsive [2Fe2S] cluster. Mol. Cell 2020, 78, 31–41.e35. [Google Scholar] [CrossRef]
- Bruening, W.; Prowse, A.H.; Schultz, D.C.; Holgado-Madruga, M.; Wong, A.; Godwin, A.K. Expression of OVCA1, a candidate tumor suppressor, is reduced in tumors and inhibits growth of ovarian cancer cells. Cancer Res. 1999, 59, 4973–4983. [Google Scholar]
- Zhang, Y.; Zhu, X.; Torelli, A.T.; Lee, M.; Dzikovski, B.; Koralewski, R.M.; Wang, E.; Freed, J.; Krebs, C.; Ealick, S.E. Diphthamide biosynthesis requires an organic radical generated by an iron–sulphur enzyme. Nature 2010, 465, 891–896. [Google Scholar] [CrossRef]
- Kwon, Y.T.; Kashina, A.S.; Varshavsky, A. Alternative splicing results in differential expression, activity, and localization of the two forms of arginyl-tRNA-protein transferase, a component of the N-end rule pathway. Mol. Cell. Biol. 1999, 19, 182–193. [Google Scholar] [CrossRef] [PubMed]
- Van, V.; Brown, J.B.; O’Shea, C.R.; Rosenbach, H.; Mohamed, I.; Ejimogu, N.-E.; Bui, T.S.; Szalai, V.A.; Chacón, K.N.; Span, I. Iron-sulfur clusters are involved in post-translational arginylation. Nat. Commun. 2023, 14, 458. [Google Scholar] [CrossRef] [PubMed]
- Du, Z.; Zhou, X.; Lai, Y.; Xu, J.; Zhang, Y.; Zhou, S.; Feng, Z.; Yu, L.; Tang, Y.; Wang, W. Structure of the human respiratory complex II. Proc. Natl. Acad. Sci. USA 2023, 120, e2216713120. [Google Scholar] [CrossRef]
- Hao, Z.; Li, X.; Qiao, T.; Du, R.; Zhang, G.; Fan, D. Subcellular localization of CIAPIN1. J. Histochem. Cytochem. 2006, 54, 1437–1444. [Google Scholar] [CrossRef]
- Banci, L.; Bertini, I.; Ciofi-Baffoni, S.; Boscaro, F.; Chatzi, A.; Mikolajczyk, M.; Tokatlidis, K.; Winkelmann, J. Anamorsin is a [2Fe-2S] cluster-containing substrate of the Mia40-dependent mitochondrial protein trapping machinery. Chem. Biol. 2011, 18, 794–804. [Google Scholar] [CrossRef]
- Zhang, Y.; Yang, C.; Dancis, A.; Nakamaru-Ogiso, E. EPR studies of wild type and mutant Dre2 identify essential [2Fe–-2S] and [4Fe–-4S] clusters and their cysteine ligands. J. Biochem. 2016, 161, 67–78. [Google Scholar] [CrossRef]
- Willems, P.; Wanschers, B.F.; Esseling, J.; Szklarczyk, R.; Kudla, U.; Duarte, I.; Forkink, M.; Nooteboom, M.; Swarts, H.; Gloerich, J. BOLA1 is an aerobic protein that prevents mitochondrial morphology changes induced by glutathione depletion. Antioxid. Redox Signal. 2013, 18, 129–138. [Google Scholar] [CrossRef]
- Li, H.; Mapolelo, D.T.; Randeniya, S.; Johnson, M.K.; Outten, C.E. Human glutaredoxin 3 forms [2Fe-2S]-bridged complexes with human BolA2. Biochemistry 2012, 51, 1687–1696. [Google Scholar] [CrossRef]
- Frey, A.G.; Palenchar, D.J.; Wildemann, J.D.; Philpott, C.C. A Glutaredoxin· BolA Complex Serves as an Iron-Sulfur Cluster Chaperone for the Cytosolic Cluster Assembly Machinery. J. Biol. Chem. 2016, 291, 22344–22356. [Google Scholar] [CrossRef]
- Pandya, P.; Braiman, A.; Isakov, N. PICOT (GLRX3) is a positive regulator of stress-induced DNA-damage response. Cell. Signal. 2019, 62, 109340. [Google Scholar] [CrossRef]
- Okuno, T.; Yamabayashi, H.; Kogure, K. Comparison of intracellular localization of Nubp1 and Nubp2 using GFP fusion proteins. Mol. Biol. Rep. 2010, 37, 1165–1168. [Google Scholar] [CrossRef] [PubMed]
- Netz, D.J.; Pierik, A.J.; Stümpfig, M.; Mühlenhoff, U.; Lill, R. The Cfd1–Nbp35 complex acts as a scaffold for iron-sulfur protein assembly in the yeast cytosol. Nat. Chem. Biol. 2007, 3, 278–286. [Google Scholar] [CrossRef] [PubMed]
- Stehling, O.; Netz, D.J.; Niggemeyer, B.; Rösser, R.; Eisenstein, R.S.; Puccio, H.; Pierik, A.J.; Lill, R. Human Nbp35 is essential for both cytosolic iron-sulfur protein assembly and iron homeostasis. Mol. Cell. Biol. 2008, 28, 5517–5528. [Google Scholar] [CrossRef] [PubMed]
- Tong, W.-H.; Jameson, G.N.; Huynh, B.H.; Rouault, T.A. Subcellular compartmentalization of human Nfu, an iron–sulfur cluster scaffold protein, and its ability to assemble a [4Fe–4S] cluster. Proc. Natl. Acad. Sci. USA 2003, 100, 9762–9767. [Google Scholar] [CrossRef]
- Cai, K.; Liu, G.; Frederick, R.O.; Xiao, R.; Montelione, G.T.; Markley, J.L. Structural/functional properties of human NFU1, an intermediate [4Fe-4S] carrier in human mitochondrial iron-sulfur cluster biogenesis. Structure 2016, 24, 2080–2091. [Google Scholar] [CrossRef]
- Land, T.; Rouault, T.A. Targeting of a human iron–sulfur cluster assembly enzyme, nifs, to different subcellular compartments is regulated through alternative AUG utilization. Mol. Cell 1998, 2, 807–815. [Google Scholar] [CrossRef]
- Fox, N.G.; Yu, X.; Feng, X.; Bailey, H.J.; Martelli, A.; Nabhan, J.F.; Strain-Damerell, C.; Bulawa, C.; Yue, W.W.; Han, S. Structure of the human frataxin-bound iron-sulfur cluster assembly complex provides insight into its activation mechanism. Nat. Commun. 2019, 10, 2210. [Google Scholar] [CrossRef]
- Marelja, Z.; Stöcklein, W.; Nimtz, M.; Leimkühler, S. A novel role for human Nfs1 in the cytoplasm: Nfs1 acts as a sulfur donor for MOCS3, a protein involved in molybdenum cofactor biosynthesis. J. Biol. Chem. 2008, 283, 25178–25185. [Google Scholar] [CrossRef]
- Lundberg, M.; Johansson, C.; Chandra, J.; Enoksson, M.; Jacobsson, G.; Ljung, J.; Johansson, M.; Holmgren, A. Cloning and expression of a novel human glutaredoxin (Grx2) with mitochondrial and nuclear isoforms. J. Biol. Chem. 2001, 276, 26269–26275. [Google Scholar] [CrossRef]
- Lillig, C.H.; Berndt, C.; Vergnolle, O.; Lönn, M.E.; Hudemann, C.; Bill, E.; Holmgren, A. Characterization of human glutaredoxin 2 as iron–sulfur protein: A possible role as redox sensor. Proc. Natl. Acad. Sci. USA 2005, 102, 8168–8173. [Google Scholar] [CrossRef]
- Kispal, G.; Csere, P.; Prohl, C.; Lill, R. The mitochondrial proteins Atm1p and Nfs1p are essential for biogenesis of cytosolic Fe/S proteins. EMBO J. 1999, 18, 3981–3989. [Google Scholar] [CrossRef] [PubMed]
- Netz, D.J.; Mascarenhas, J.; Stehling, O.; Pierik, A.J.; Lill, R. Maturation of cytosolic and nuclear iron–sulfur proteins. Trends Cell Biol. 2014, 24, 303–312. [Google Scholar] [CrossRef] [PubMed]
- Song, D.; Lee, F.S. A role for IOP1 in mammalian cytosolic iron-sulfur protein biogenesis. J. Biol. Chem. 2008, 283, 9231–9238. [Google Scholar] [CrossRef] [PubMed]
- Netz, D.J.; Pierik, A.J.; Stümpfig, M.; Bill, E.; Sharma, A.K.; Pallesen, L.J.; Walden, W.E.; Lill, R. A bridging [4Fe-4S] cluster and nucleotide binding are essential for function of the Cfd1-Nbp35 complex as a scaffold in iron-sulfur protein maturation. J. Biol. Chem. 2012, 287, 12365–12378. [Google Scholar] [CrossRef]
- Stehling, O.; Vashisht, A.A.; Mascarenhas, J.; Jonsson, Z.O.; Sharma, T.; Netz, D.J.; Pierik, A.J.; Wohlschlegel, J.A.; Lill, R. MMS19 assembles iron-sulfur proteins required for DNA metabolism and genomic integrity. Science 2012, 337, 195–199. [Google Scholar] [CrossRef]
- Gari, K.; León Ortiz, A.M.; Borel, V.; Flynn, H.; Skehel, J.M.; Boulton, S.J. MMS19 links cytoplasmic iron-sulfur cluster assembly to DNA metabolism. Science 2012, 337, 243–245. [Google Scholar] [CrossRef]
- Balk, J.; Pierik, A.J.; Netz, D.J.A.; Mühlenhoff, U.; Lill, R. The hydrogenase-like Nar1p is essential for maturation of cytosolic and nuclear iron–sulphur proteins. EMBO J. 2004, 23, 2105–2115. [Google Scholar] [CrossRef]
- SantaMaria, A.M.; Rouault, T.A. Regulatory and Sensing Iron–Sulfur Clusters: New Insights and Unanswered Questions. Inorganics 2024, 12, 101. [Google Scholar] [CrossRef]
- Querci, L.; Piccioli, M.; Ciofi-Baffoni, S.; Banci, L. Structural aspects of iron-sulfur protein biogenesis: An NMR view. Biochim. Et. Biophys. Acta (BBA)-Mol. Cell Res. 2024, 1871, 119786. [Google Scholar] [CrossRef]
- Maio, N.; Singh, A.; Uhrigshardt, H.; Saxena, N.; Tong, W.-H.; Rouault, T.A. Cochaperone binding to LYR motifs confers specificity of iron sulfur cluster delivery. Cell Metab. 2014, 19, 445–457. [Google Scholar] [CrossRef]
- Li, K.; Tong, W.-H.; Hughes, R.M.; Rouault, T.A. Roles of the mammalian cytosolic cysteine desulfurase, ISCS, and scaffold protein, ISCU, in iron-sulfur cluster assembly. J. Biol. Chem. 2006, 281, 12344–12351. [Google Scholar] [CrossRef]
- Kim, K.S.; Maio, N.; Singh, A.; Rouault, T.A. Cytosolic HSC20 integrates de novo iron–sulfur cluster biogenesis with the CIAO1-mediated transfer to recipients. Hum. Mol. Genet. 2018, 27, 837–852. [Google Scholar] [CrossRef] [PubMed]
- Shi, H.; Bencze, K.Z.; Stemmler, T.L.; Philpott, C.C. A cytosolic iron chaperone that delivers iron to ferritin. Science 2008, 320, 1207–1210. [Google Scholar] [CrossRef] [PubMed]
- Philpott, C.C. Coming into view: Eukaryotic iron chaperones and intracellular iron delivery. J. Biol. Chem. 2012, 287, 13518–13523. [Google Scholar] [CrossRef] [PubMed]
- Philpott, C.C.; Ryu, M.-S.; Frey, A.; Patel, S. Cytosolic iron chaperones: Proteins delivering iron cofactors in the cytosol of mammalian cells. J. Biol. Chem. 2017, 292, 12764–12771. [Google Scholar] [CrossRef]
- Philpott, C.C.; Patel, S.J.; Protchenko, O. Management versus miscues in the cytosolic labile iron pool: The varied functions of iron chaperones. Biochim. Et Biophys. Acta (BBA)-Mol. Cell Res. 2020, 1867, 118830. [Google Scholar] [CrossRef]
- Patel, S.J.; Frey, A.G.; Palenchar, D.J.; Achar, S.; Bullough, K.Z.; Vashisht, A.; Wohlschlegel, J.A.; Philpott, C.C. A PCBP1–BolA2 chaperone complex delivers iron for cytosolic [2Fe–2S] cluster assembly. Nat. Chem. Biol. 2019, 15, 872–881. [Google Scholar] [CrossRef]
- Adamec, J.; Rusnak, F.; Owen, W.G.; Naylor, S.; Benson, L.M.; Gacy, A.M.; Isaya, G. Iron-dependent self-assembly of recombinant yeast frataxin: Implications for Friedreich ataxia. Am. J. Hum. Genet. 2000, 67, 549–562. [Google Scholar] [CrossRef]
- Bulteau, A.-L.; O’Neill, H.A.; Kennedy, M.C.; Ikeda-Saito, M.; Isaya, G.; Szweda, L.I. Frataxin acts as an iron chaperone protein to modulate mitochondrial aconitase activity. Science 2004, 305, 242–245. [Google Scholar] [CrossRef]
- Cavadini, P.; O’Neill, H.A.; Benada, O.; Isaya, G. Assembly and iron-binding properties of human frataxin, the protein deficient in Friedreich ataxia. Hum. Mol. Genet. 2002, 11, 217–227. [Google Scholar] [CrossRef]
- Bencze, K.Z.; Kondapalli, K.C.; Cook, J.D.; McMahon, S.; Millán-Pacheco, C.; Pastor, N.; Stemmler, T.L. The structure and function of frataxin. Crit. Rev. Biochem. Mol. Biol. 2006, 41, 269–291. [Google Scholar] [CrossRef] [PubMed]
- Parent, A.; Elduque, X.; Cornu, D.; Belot, L.; Le Caer, J.-P.; Grandas, A.; Toledano, M.B.; D’autréaux, B. Mammalian frataxin directly enhances sulfur transfer of NFS1 persulfide to both ISCU and free thiols. Nat. Commun. 2015, 6, 5686. [Google Scholar] [CrossRef] [PubMed]
- Bridwell-Rabb, J.; Fox, N.G.; Tsai, C.-L.; Winn, A.M.; Barondeau, D.P. Human frataxin activates Fe–S cluster biosynthesis by facilitating sulfur transfer chemistry. Biochemistry 2014, 53, 4904–4913. [Google Scholar] [CrossRef]
- Balk, J.; Aguilar Netz, D.J.; Tepper, K.; Pierik, A.J.; Lill, R. The essential WD40 protein Cia1 is involved in a late step of cytosolic and nuclear iron-sulfur protein assembly. Mol. Cell. Biol. 2005, 25, 10833–10841. [Google Scholar] [CrossRef]
- Stehling, O.; Mascarenhas, J.; Vashisht, A.A.; Sheftel, A.D.; Niggemeyer, B.; Rösser, R.; Pierik, A.J.; Wohlschlegel, J.A.; Lill, R. Human CIA2A-FAM96A and CIA2B-FAM96B integrate iron homeostasis and maturation of different subsets of cytosolic-nuclear iron-sulfur proteins. Cell Metab. 2013, 18, 187–198. [Google Scholar] [CrossRef]
- Marquez, M.D.; Greth, C.; Buzuk, A.; Liu, Y.; Blinn, C.M.; Beller, S.; Leiskau, L.; Hushka, A.; Wu, K.; Nur, K. Cytosolic iron–sulfur protein assembly system identifies clients by a C-terminal tripeptide. Proc. Natl. Acad. Sci. USA 2023, 120, e2311057120. [Google Scholar] [CrossRef]
- Paul, V.D.; Mühlenhoff, U.; Stümpfig, M.; Seebacher, J.; Kugler, K.G.; Renicke, C.; Taxis, C.; Gavin, A.-C.; Pierik, A.J.; Lill, R. The deca-GX3 proteins Yae1-Lto1 function as adaptors recruiting the ABC protein Rli1 for iron-sulfur cluster insertion. Elife 2015, 4, e08231. [Google Scholar] [CrossRef]
- Odermatt, D.C.; Gari, K. The CIA Targeting Complex Is Highly Regulated and Provides Two Distinct Binding Sites for Client Iron-Sulfur Proteins. Cell Rep. 2017, 18, 1434–1443. [Google Scholar] [CrossRef]
- Kassube, S.A.; Thomä, N.H. Structural insights into Fe–S protein biogenesis by the CIA targeting complex. Nat. Struct. Mol. Biol. 2020, 27, 735–742. [Google Scholar] [CrossRef]
- Alhebshi, A.; Sideri, T.C.; Holland, S.L.; Avery, S.V. The essential iron-sulfur protein Rli1 is an important target accounting for inhibition of cell growth by reactive oxygen species. Mol. Biol. Cell 2012, 23, 3582–3590. [Google Scholar] [CrossRef]
- Honarmand Ebrahimi, K.; Ciofi-Baffoni, S.; Hagedoorn, P.-L.; Nicolet, Y.; Le Brun, N.E.; Hagen, W.R.; Armstrong, F.A. Iron–sulfur clusters as inhibitors and catalysts of viral replication. Nat. Chem. 2022, 14, 253–266. [Google Scholar] [CrossRef]
- Jang, S.; Imlay, J.A. Micromolar intracellular hydrogen peroxide disrupts metabolism by damaging iron-sulfur enzymes. J. Biol. Chem. 2007, 282, 929–937. [Google Scholar] [CrossRef]
- Mettert, E.L.; Kiley, P.J. Fe–S proteins that regulate gene expression. Biochim. Et Biophys. Acta (BBA)-Mol. Cell Res. 2015, 1853, 1284–1293. [Google Scholar] [CrossRef] [PubMed]
- Imlay, J.A. Iron-sulphur clusters and the problem with oxygen. Mol. Microbiol. 2006, 59, 1073–1082. [Google Scholar] [CrossRef]
- Mettert, E.L.; Kiley, P.J. How is Fe-S cluster formation regulated? Annu. Rev. Microbiol. 2015, 69, 505–526. [Google Scholar] [CrossRef]
- Lauble, H.; Kennedy, M.; Beinert, H.; Stout, C. Crystal structures of aconitase with isocitrate and nitroisocitrate bound. Biochemistry 1992, 31, 2735–2748. [Google Scholar] [CrossRef]
- Federico, G.; Carrillo, F.; Dapporto, F.; Chiariello, M.; Santoro, M.; Bellelli, R.; Carlomagno, F. NCOA4 links iron bioavailability to DNA metabolism. Cell Rep. 2022, 40, 111207. [Google Scholar] [CrossRef]
- Mayank, A.K.; Pandey, V.; Vashisht, A.A.; Barshop, W.D.; Rayatpisheh, S.; Sharma, T.; Haque, T.; Powers, D.N.; Wohlschlegel, J.A. An oxygen-dependent interaction between FBXL5 and the CIA-targeting complex regulates iron homeostasis. Mol. Cell 2019, 75, 382–393.e385. [Google Scholar] [CrossRef]
- van Wietmarschen, N.; Moradian, A.; Morin, G.B.; Lansdorp, P.M.; Uringa, E.-J. The mammalian proteins MMS19, MIP18, and ANT2 are involved in cytoplasmic iron-sulfur cluster protein assembly. J. Biol. Chem. 2012, 287, 43351–43358. [Google Scholar] [CrossRef]
- Puig, S.; Ramos-Alonso, L.; Romero, A.; Martínez-Pastor, M. The elemental role of iron in DNA synthesis and repair. Metallomics 2017, 9, 1483–1500. [Google Scholar] [CrossRef]
- Troadec, M.-B.; Loréal, O.; Brissot, P. The interaction of iron and the genome: For better and for worse. Mutat. Res./Rev. Mutat. Res. 2017, 774, 25–32. [Google Scholar] [CrossRef] [PubMed]
- Veatch, J.R.; McMurray, M.A.; Nelson, Z.W.; Gottschling, D.E. Mitochondrial dysfunction leads to nuclear genome instability via an iron-sulfur cluster defect. Cell 2009, 137, 1247–1258. [Google Scholar] [CrossRef] [PubMed]
- Shi, R.; Hou, W.; Wang, Z.-Q.; Xu, X. Biogenesis of iron–sulfur clusters and their role in DNA metabolism. Front. Cell Dev. Biol. 2021, 9, 735678. [Google Scholar] [CrossRef]
- Loeb, L.A.; Monnat, R.J., Jr. DNA polymerases and human disease. Nat. Rev. Genet. 2008, 9, 594–604. [Google Scholar] [CrossRef]
- Ter Beek, J.; Parkash, V.; Bylund, G.O.; Osterman, P.; Sauer-Eriksson, A.E.; Johansson, E. Structural evidence for an essential Fe–S cluster in the catalytic core domain of DNA polymerase ϵ. Nucleic Acids Res. 2019, 47, 5712–5722. [Google Scholar] [CrossRef]
- Jain, R.; Vanamee, E.S.; Dzikovski, B.G.; Buku, A.; Johnson, R.E.; Prakash, L.; Prakash, S.; Aggarwal, A.K. An iron–sulfur cluster in the polymerase domain of yeast DNA polymerase ε. J. Mol. Biol. 2014, 426, 301–308. [Google Scholar] [CrossRef]
- Baranovskiy, A.G.; Lada, A.G.; Siebler, H.M.; Zhang, Y.; Pavlov, Y.I.; Tahirov, T.H. DNA polymerase δ and ζ switch by sharing accessory subunits of DNA polymerase δ. J. Biol. Chem. 2012, 287, 17281–17287. [Google Scholar] [CrossRef]
- Lisova, A.E.; Baranovskiy, A.G.; Morstadt, L.M.; Babayeva, N.D.; Stepchenkova, E.I.; Tahirov, T.H. The iron-sulfur cluster is essential for DNA binding by human DNA polymerase ε. Sci. Rep. 2022, 12, 17436. [Google Scholar] [CrossRef]
- Bartels, P.L.; Stodola, J.L.; Burgers, P.M.; Barton, J.K. A redox role for the [4Fe4S] cluster of yeast DNA polymerase δ. J. Am. Chem. Soc. 2017, 139, 18339–18348. [Google Scholar] [CrossRef]
- Baranovskiy, A.G.; Babayeva, N.D.; Zhang, Y.; Gu, J.; Suwa, Y.; Pavlov, Y.I.; Tahirov, T.H. Mechanism of concerted RNA-DNA primer synthesis by the human primosome. J. Biol. Chem. 2016, 291, 10006–10020. [Google Scholar] [CrossRef]
- Suwa, Y.; Gu, J.; Baranovskiy, A.G.; Babayeva, N.D.; Pavlov, Y.I.; Tahirov, T.H. Crystal structure of the human Pol α B subunit in complex with the C-terminal domain of the catalytic subunit. J. Biol. Chem. 2015, 290, 14328–14337. [Google Scholar] [CrossRef] [PubMed]
- Klinge, S.; Núñez-Ramírez, R.; Llorca, O.; Pellegrini, L. 3D architecture of DNA Pol α reveals the functional core of multi-subunit replicative polymerases. EMBO J. 2009, 28, 1978–1987. [Google Scholar] [CrossRef] [PubMed]
- Zhang, Y.; Baranovskiy, A.G.; Tahirov, T.H.; Pavlov, Y.I. The C-terminal domain of the DNA polymerase catalytic subunit regulates the primase and polymerase activities of the human DNA polymerase α-primase complex. J. Biol. Chem. 2014, 289, 22021–22034. [Google Scholar] [CrossRef]
- Baranovskiy, A.G.; Tahirov, T.H. Elaborated action of the human primosome. Genes 2017, 8, 62. [Google Scholar] [CrossRef]
- Pritts, J.D.; Michel, S.L. Fe-S clusters masquerading as zinc finger proteins. J. Inorg. Biochem. 2022, 230, 111756. [Google Scholar] [CrossRef]
- Maio, N.; Raza, M.K.; Li, Y.; Zhang, D.-L.; Bollinger, J.M., Jr.; Krebs, C.; Rouault, T.A. An iron–sulfur cluster in the zinc-binding domain of the SARS-CoV-2 helicase modulates its RNA-binding and-unwinding activities. Proc. Natl. Acad. Sci. USA 2023, 120, e2303860120. [Google Scholar] [CrossRef]
- Maio, N.; Lafont, B.A.; Sil, D.; Li, Y.; Bollinger, J.M., Jr.; Krebs, C.; Pierson, T.C.; Linehan, W.M.; Rouault, T.A. Fe-S cofactors in the SARS-CoV-2 RNA-dependent RNA polymerase are potential antiviral targets. Science 2021, 373, 236–241. [Google Scholar] [CrossRef]
- Conlan, A.R.; Axelrod, H.L.; Cohen, A.E.; Abresch, E.C.; Zuris, J.; Yee, D.; Nechushtai, R.; Jennings, P.A.; Paddock, M.L. Crystal structure of Miner1: The redox-active 2Fe-2S protein causative in Wolfram Syndrome 2. J. Mol. Biol. 2009, 392, 143–153. [Google Scholar] [CrossRef]
- Cutone, A.; Howes, B.D.; Miele, A.E.; Miele, R.; Giorgi, A.; Battistoni, A.; Smulevich, G.; Musci, G.; Di Patti, M.C.B. Pichia pastoris Fep1 is a [2Fe-2S] protein with a Zn finger that displays an unusual oxygen-dependent role in cluster binding. Sci. Rep. 2016, 6, 31872. [Google Scholar] [CrossRef]
- Liu, L.; Huang, M. Essential role of the iron-sulfur cluster binding domain of the primase regulatory subunit Pri2 in DNA replication initiation. Protein Cell 2015, 6, 194–210. [Google Scholar] [CrossRef]
- Baranovskiy, A.G.; Zhang, Y.; Suwa, Y.; Babayeva, N.D.; Gu, J.; Pavlov, Y.I.; Tahirov, T.H. Crystal structure of the human primase. J. Biol. Chem. 2015, 290, 5635–5646. [Google Scholar] [CrossRef] [PubMed]
- Sauguet, L.; Klinge, S.; Perera, R.L.; Maman, J.D.; Pellegrini, L. Shared active site architecture between the large subunit of eukaryotic primase and DNA photolyase. PLoS ONE 2010, 5, e10083. [Google Scholar] [CrossRef] [PubMed]
- Agarkar, V.B.; Babayeva, N.D.; Pavlov, Y.I.; Tahirov, T.H. Crystal structure of the C-terminal domain of human DNA primase large subunit: Implications for the mechanism of the primase-polymerase α switch. Cell Cycle 2011, 10, 926–931. [Google Scholar] [CrossRef] [PubMed]
- O’Brien, E.; Holt, M.E.; Thompson, M.K.; Salay, L.E.; Ehlinger, A.C.; Chazin, W.J.; Barton, J.K. The [4Fe4S] cluster of human DNA primase functions as a redox switch using DNA charge transport. Science 2017, 355, eaag1789. [Google Scholar] [CrossRef]
- O’Brien, E.; Salay, L.E.; Epum, E.A.; Friedman, K.L.; Chazin, W.J.; Barton, J.K. Yeast require redox switching in DNA primase. Proc. Natl. Acad. Sci. USA 2018, 115, 13186–13191. [Google Scholar] [CrossRef]
- Amin, M.; Brooks, B.R. The oxidation of the [4Fe-4S] cluster of DNA primase alters the binding energies with DNA and RNA primers. Biophys. J. 2024, 123, 1648–1653. [Google Scholar] [CrossRef]
- O’Brien, E.; Holt, M.E.; Salay, L.E.; Chazin, W.J.; Barton, J.K. Substrate binding regulates redox signaling in human DNA primase. J. Am. Chem. Soc. 2018, 140, 17153–17162. [Google Scholar] [CrossRef]
- Mancias, J.D.; Wang, X.; Gygi, S.P.; Harper, J.W.; Kimmelman, A.C. Quantitative proteomics identifies NCOA4 as the cargo receptor mediating ferritinophagy. Nature 2014, 509, 105–109. [Google Scholar] [CrossRef]
- Dowdle, W.E.; Nyfeler, B.; Nagel, J.; Elling, R.A.; Liu, S.; Triantafellow, E.; Menon, S.; Wang, Z.; Honda, A.; Pardee, G. Selective VPS34 inhibitor blocks autophagy and uncovers a role for NCOA4 in ferritin degradation and iron homeostasis in vivo. Nat. Cell Biol. 2014, 16, 1069–1079. [Google Scholar] [CrossRef]
- Singleton, M.R.; Dillingham, M.S.; Wigley, D.B. Structure and mechanism of helicases and nucleic acid translocases. Annu. Rev. Biochem. 2007, 76, 23–50. [Google Scholar] [CrossRef]
- Wu, Y.; Brosh, R.M., Jr. DNA helicase and helicase–nuclease enzymes with a conserved iron–sulfur cluster. Nucleic Acids Res. 2012, 40, 4247–4260. [Google Scholar] [CrossRef] [PubMed]
- Yeeles, J.T.; Cammack, R.; Dillingham, M.S. An iron-sulfur cluster is essential for the binding of broken DNA by AddAB-type helicase-nucleases. J. Biol. Chem. 2009, 284, 7746–7755. [Google Scholar] [CrossRef] [PubMed]
- Zheng, L.; Zhou, M.; Guo, Z.; Lu, H.; Qian, L.; Dai, H.; Qiu, J.; Yakubovskaya, E.; Bogenhagen, D.F.; Demple, B. Human DNA2 is a mitochondrial nuclease/helicase for efficient processing of DNA replication and repair intermediates. Mol. Cell 2008, 32, 325–336. [Google Scholar] [CrossRef] [PubMed]
- Estep, K.N.; Brosh, R.M., Jr. RecQ and Fe–S helicases have unique roles in DNA metabolism dictated by their unwinding directionality, substrate specificity, and protein interactions. Biochem. Soc. Trans. 2018, 46, 77–95. [Google Scholar] [CrossRef]
- Matsuzaki, K.; Borel, V.; Adelman, C.A.; Schindler, D.; Boulton, S.J. FANCJ suppresses microsatellite instability and lymphomagenesis independent of the Fanconi anemia pathway. Genes Dev. 2015, 29, 2532–2546. [Google Scholar] [CrossRef]
- Capo-Chichi, J.M.; Bharti, S.K.; Sommers, J.A.; Yammine, T.; Chouery, E.; Patry, L.; Rouleau, G.A.; Samuels, M.E.; Hamdan, F.F.; Michaud, J.L. Identification and biochemical characterization of a novel mutation in DDX11 causing W arsaw breakage syndrome. Hum. Mutat. 2013, 34, 103–107. [Google Scholar] [CrossRef]
- Deng, Z.; Glousker, G.; Molczan, A.; Fox, A.J.; Lamm, N.; Dheekollu, J.; Weizman, O.-E.; Schertzer, M.; Wang, Z.; Vladimirova, O. Inherited mutations in the helicase RTEL1 cause telomere dysfunction and Hoyeraal–Hreidarsson syndrome. Proc. Natl. Acad. Sci. USA 2013, 110, E3408–E3416. [Google Scholar] [CrossRef]
- Falquet, B.; Ölmezer, G.; Enkner, F.; Klein, D.; Challa, K.; Appanah, R.; Gasser, S.M.; Rass, U. Disease-associated DNA2 nuclease–helicase protects cells from lethal chromosome under-replication. Nucleic Acids Res. 2020, 48, 7265–7278. [Google Scholar] [CrossRef]
- Sommers, J.A.; Rawtani, N.; Gupta, R.; Bugreev, D.V.; Mazin, A.V.; Cantor, S.B.; Brosh, R.M. FANCJ uses its motor ATPase to destabilize protein-DNA complexes, unwind triplexes, and inhibit RAD51 strand exchange. J. Biol. Chem. 2009, 284, 7505–7517. [Google Scholar] [CrossRef]
- Levran, O.; Attwooll, C.; Henry, R.T.; Milton, K.L.; Neveling, K.; Rio, P.; Batish, S.D.; Kalb, R.; Velleuer, E.; Barral, S. The BRCA1-interacting helicase BRIP1 is deficient in Fanconi anemia. Nat. Genet. 2005, 37, 931–933. [Google Scholar] [CrossRef]
- Odermatt, D.C.; Lee, W.T.C.; Wild, S.; Jozwiakowski, S.K.; Rothenberg, E.; Gari, K. Cancer-associated mutations in the iron-sulfur domain of FANCJ affect G-quadruplex metabolism. PLoS Genet. 2020, 16, e1008740. [Google Scholar] [CrossRef] [PubMed]
- Bharti, S.K.; Sommers, J.A.; George, F.; Kuper, J.; Hamon, F.; Shin-ya, K.; Teulade-Fichou, M.-P.; Kisker, C.; Brosh, R.M. Specialization among iron-sulfur cluster helicases to resolve G-quadruplex DNA structures that threaten genomic stability. J. Biol. Chem. 2013, 288, 28217–28229. [Google Scholar] [CrossRef] [PubMed]
- Wulfridge, P.; Sarma, K. Intertwining roles of R-loops and G-quadruplexes in DNA repair, transcription and genome organization. Nat. Cell Biol. 2024, 26, 1025–1036. [Google Scholar] [CrossRef] [PubMed]
- Hänsel-Hertsch, R.; Di Antonio, M.; Balasubramanian, S. DNA G-quadruplexes in the human genome: Detection, functions and therapeutic potential. Nat. Rev. Mol. Cell Biol. 2017, 18, 279–284. [Google Scholar] [CrossRef]
- Wu, Y.; Shin-ya, K.; Brosh, R.M., Jr. FANCJ helicase defective in Fanconia anemia and breast cancer unwinds G-quadruplex DNA to defend genomic stability. Mol. Cell. Biol. 2008, 28, 4116–4128. [Google Scholar] [CrossRef]
- London, T.B.; Barber, L.J.; Mosedale, G.; Kelly, G.P.; Balasubramanian, S.; Hickson, I.D.; Boulton, S.J.; Hiom, K. FANCJ is a structure-specific DNA helicase associated with the maintenance of genomic G/C tracts. J. Biol. Chem. 2008, 283, 36132–36139. [Google Scholar] [CrossRef]
- Wu, C.G.; Spies, M. G-quadruplex recognition and remodeling by the FANCJ helicase. Nucleic Acids Res. 2016, 44, 8742–8753. [Google Scholar] [CrossRef]
- Vannier, J.-B.; Sandhu, S.; Petalcorin, M.I.; Wu, X.; Nabi, Z.; Ding, H.; Boulton, S.J. RTEL1 is a replisome-associated helicase that promotes telomere and genome-wide replication. Science 2013, 342, 239–242. [Google Scholar] [CrossRef]
- Uringa, E.-J.; Youds, J.L.; Lisaingo, K.; Lansdorp, P.M.; Boulton, S.J. RTEL1: An essential helicase for telomere maintenance and the regulation of homologous recombination. Nucleic Acids Res. 2011, 39, 1647–1655. [Google Scholar] [CrossRef]
- Wu, W.; Bhowmick, R.; Vogel, I.; Özer, Ö.; Ghisays, F.; Thakur, R.S.; Sanchez de Leon, E.; Richter, P.H.; Ren, L.; Petrini, J.H. RTEL1 suppresses G-quadruplex-associated R-loops at difficult-to-replicate loci in the human genome. Nat. Struct. Mol. Biol. 2020, 27, 424–437. [Google Scholar] [CrossRef]
- Ghisays, F.; Garzia, A.; Wang, H.; Canasto-Chibuque, C.; Hohl, M.; Savage, S.A.; Tuschl, T.; Petrini, J.H. RTEL1 influences the abundance and localization of TERRA RNA. Nat. Commun. 2021, 12, 3016. [Google Scholar] [CrossRef] [PubMed]
- Wu, Y.; Sommers, J.A.; Khan, I.; de Winter, J.P.; Brosh, R.M. Biochemical characterization of Warsaw breakage syndrome helicase. J. Biol. Chem. 2012, 287, 1007–1021. [Google Scholar] [CrossRef] [PubMed]
- van Schie, J.J.; Faramarz, A.; Balk, J.A.; Stewart, G.S.; Cantelli, E.; Oostra, A.B.; Rooimans, M.A.; Parish, J.L.; de Almeida Estéves, C.; Dumic, K. Warsaw Breakage Syndrome associated DDX11 helicase resolves G-quadruplex structures to support sister chromatid cohesion. Nat. Commun. 2020, 11, 4287. [Google Scholar] [CrossRef] [PubMed]
- Gray, L.T.; Vallur, A.C.; Eddy, J.; Maizels, N. G quadruplexes are genomewide targets of transcriptional helicases XPB and XPD. Nat. Chem. Biol. 2014, 10, 313–318. [Google Scholar] [CrossRef] [PubMed]
- Kou, H.; Zhou, Y.; Gorospe, R.C.; Wang, Z. Mms19 protein functions in nucleotide excision repair by sustaining an adequate cellular concentration of the TFIIH component Rad3. Proc. Natl. Acad. Sci. USA 2008, 105, 15714–15719. [Google Scholar] [CrossRef]
- Seroz, T.; Winkler, G.S.; Auriol, J.; Verhage, R.A.; Vermeulen, W.; Smit, B.; Brouwer, J.; Eker, A.P.; Weeda, G.; Egly, J.-M. Cloning of a human homolog of the yeast nucleotide excision repair gene MMS19 and interaction with transcription repair factor TFIIH via the XPB and XPD helicases. Nucleic Acids Res. 2000, 28, 4506–4513. [Google Scholar] [CrossRef]
- Ding, H.; Schertzer, M.; Wu, X.; Gertsenstein, M.; Selig, S.; Kammori, M.; Pourvali, R.; Poon, S.; Vulto, I.; Chavez, E. Regulation of murine telomere length by Rtel: An essential gene encoding a helicase-like protein. Cell 2004, 117, 873–886. [Google Scholar] [CrossRef]
- Askree, S.H.; Yehuda, T.; Smolikov, S.; Gurevich, R.; Hawk, J.; Coker, C.; Krauskopf, A.; Kupiec, M.; McEachern, M.J. A genome-wide screen for Saccharomyces cerevisiae deletion mutants that affect telomere length. Proc. Natl. Acad. Sci. USA 2004, 101, 8658–8663. [Google Scholar] [CrossRef]
- Friedberg, E.C.; Walker, G.C.; Siede, W.; Wood, R.D. DNA Repair and Mutagenesis; American Society for Microbiology Press: Washington, DC, USA, 2005. [Google Scholar]
- Knijnenburg, T.A.; Wang, L.; Zimmermann, M.T.; Chambwe, N.; Gao, G.F.; Cherniack, A.D.; Fan, H.; Shen, H.; Way, G.P.; Greene, C.S. Genomic and molecular landscape of DNA damage repair deficiency across the cancer genome atlas. Cell Rep. 2018, 23, 239–254.e236. [Google Scholar] [CrossRef]
- Fuss, J.O.; Tsai, C.-L.; Ishida, J.P.; Tainer, J.A. Emerging critical roles of Fe–S clusters in DNA replication and repair. Biochim. Et Biophys. Acta (BBA)-Mol. Cell Res. 2015, 1853, 1253–1271. [Google Scholar] [CrossRef]
- Cunningham, R.P.; Asahara, H.; Bank, J.F.; Scholes, C.P.; Salerno, J.C.; Surerus, K.; Munck, E.; McCracken, J.; Peisach, J.; Emptage, M.H. Endonuclease III is an iron-sulfur protein. Biochemistry 1989, 28, 4450–4455. [Google Scholar] [CrossRef] [PubMed]
- Hinks, J.A.; Evans, M.C.; De Miguel, Y.; Sartori, A.A.; Jiricny, J.; Pearl, L.H. An iron-sulfur cluster in the family 4 uracil-DNA glycosylases. J. Biol. Chem. 2002, 277, 16936–16940. [Google Scholar] [CrossRef] [PubMed]
- Parker, A.; Gu, Y.; Lu, A.-L. Purification and characterization of a mammalian homolog of Escherichia coli MutY mismatch repair protein from calf liver mitochondria. Nucleic Acids Res. 2000, 28, 3206–3215. [Google Scholar] [CrossRef] [PubMed]
- Jacobs, A.L.; Schär, P. DNA glycosylases: In DNA repair and beyond. Chromosoma 2012, 121, 1–20. [Google Scholar] [CrossRef]
- McGoldrick, J.P.; Yeh, Y.-C.; Solomon, M.; Essigmann, J.M.; Lu, A.-L. Characterization of a mammalian homolog of the Escherichia coli MutY mismatch repair protein. Mol. Cell. Biol. 1995, 15, 989–996. [Google Scholar] [CrossRef]
- Aspinwall, R.; Rothwell, D.G.; Roldan-Arjona, T.; Anselmino, C.; Ward, C.J.; Cheadle, J.P.; Sampson, J.R.; Lindahl, T.; Harris, P.C.; Hickson, I.D. Cloning and characterization of a functional human homolog of Escherichia coli endonuclease III. Proc. Natl. Acad. Sci. USA 1997, 94, 109–114. [Google Scholar] [CrossRef]
- Lukianova, O.A.; David, S.S. A role for iron–sulfur clusters in DNA repair. Curr. Opin. Chem. Biol. 2005, 9, 145–151. [Google Scholar] [CrossRef]
- Trasviña-Arenas, C.H.; Lopez-Castillo, L.M.; Sanchez-Sandoval, E.; Brieba, L.G. Dispensability of the [4Fe-4S] cluster in novel homologues of adenine glycosylase MutY. FEBS J. 2016, 283, 521–540. [Google Scholar] [CrossRef]
- Guan, Y.; Manuel, R.C.; Arvai, A.S.; Parikh, S.S.; Mol, C.D.; Miller, J.H.; Lloyd, R.S.; Tainer, J.A. MutY catalytic core, mutant and bound adenine structures define specificity for DNA repair enzyme superfamily. Nat. Struct. Biol. 1998, 5, 1058–1064. [Google Scholar] [CrossRef]
- Boal, A.K.; Yavin, E.; Lukianova, O.A.; O’Shea, V.L.; David, S.S.; Barton, J.K. DNA-bound redox activity of DNA repair glycosylases containing [4Fe-4S] clusters. Biochemistry 2005, 44, 8397–8407. [Google Scholar] [CrossRef]
- Romano, C.A.; Sontz, P.A.; Barton, J.K. Mutants of the base excision repair glycosylase, endonuclease III: DNA charge transport as a first step in lesion detection. Biochemistry 2011, 50, 6133–6145. [Google Scholar] [CrossRef] [PubMed]
- Porello, S.L.; Cannon, M.J.; David, S.S. A substrate recognition role for the [4Fe-4S] 2+ cluster of the DNA repair glycosylase MutY. Biochemistry 1998, 37, 6465–6475. [Google Scholar] [CrossRef] [PubMed]
- Bartels, P.L.; Zhou, A.; Arnold, A.R.; Nuñez, N.N.; Crespilho, F.N.; David, S.S.; Barton, J.K. Electrochemistry of the [4Fe4S] cluster in base excision repair proteins: Tuning the redox potential with DNA. Langmuir 2017, 33, 2523–2530. [Google Scholar] [CrossRef] [PubMed]
- Tse, E.C.; Zwang, T.J.; Barton, J.K. The oxidation state of [4Fe4S] clusters modulates the DNA-binding affinity of DNA repair proteins. J. Am. Chem. Soc. 2017, 139, 12784–12792. [Google Scholar] [CrossRef]
- Bruner, S.D.; Norman, D.P.; Verdine, G.L. Structural basis for recognition and repair of the endogenous mutagen 8-oxoguanine in DNA. Nature 2000, 403, 859–866. [Google Scholar] [CrossRef]
- Poor, C.B.; Wegner, S.V.; Li, H.; Dlouhy, A.C.; Schuermann, J.P.; Sanishvili, R.; Hinshaw, J.R.; Riggs-Gelasco, P.J.; Outten, C.E.; He, C. Molecular mechanism and structure of the Saccharomyces cerevisiae iron regulator Aft2. Proc. Natl. Acad. Sci. USA 2014, 111, 4043–4048. [Google Scholar] [CrossRef]
- Ueta, R.; Fujiwara, N.; Iwai, K.; Yamaguchi-Iwai, Y. Iron-induced dissociation of the Aft1p transcriptional regulator from target gene promoters is an initial event in iron-dependent gene suppression. Mol. Cell Biol. 2012, 32, 4998–5008. [Google Scholar] [CrossRef]
- Li, H.; Mapolelo, D.T.; Dingra, N.N.; Naik, S.G.; Lees, N.S.; Hoffman, B.M.; Riggs-Gelasco, P.J.; Huynh, B.H.; Johnson, M.K.; Outten, C.E. The yeast iron regulatory proteins Grx3/4 and Fra2 form heterodimeric complexes containing a [2Fe-2S] cluster with cysteinyl and histidyl ligation. Biochemistry 2009, 48, 9569–9581. [Google Scholar] [CrossRef]
- Mercier, A.; Watt, S.; Bähler, J.; Labbé, S. Key function for the CCAAT-binding factor Php4 to regulate gene expression in response to iron deficiency in fission yeast. Eukaryot. Cell 2008, 7, 493–508. [Google Scholar] [CrossRef]
- Vachon, P.; Mercier, A.; Jbel, M.; Labbé, S. The monothiol glutaredoxin Grx4 exerts an iron-dependent inhibitory effect on Php4 function. Eukaryot. Cell 2012, 11, 806–819. [Google Scholar] [CrossRef]
- Dlouhy, A.C.; Beaudoin, J.; Labbé, S.; Outten, C.E. Schizosaccharomyces pombe Grx4 regulates the transcriptional repressor Php4 via [2Fe–2S] cluster binding. Metallomics 2017, 9, 1096–1105. [Google Scholar] [CrossRef] [PubMed]
- Hati, D.; Brault, A.; Gupta, M.; Fletcher, K.; Jacques, J.F.; Labbe, S.; Outten, C.E. Iron homeostasis proteins Grx4 and Fra2 control activity of the Schizosaccharomyces pombe iron repressor Fep1 by facilitating [2Fe-2S] cluster removal. J. Biol. Chem. 2023, 299, 105419. [Google Scholar] [CrossRef] [PubMed]
- Jacques, J.F.; Mercier, A.; Brault, A.; Mourer, T.; Labbe, S. Fra2 is a co-regulator of Fep1 inhibition in response to iron starvation. PLoS ONE 2014, 9, e98959. [Google Scholar] [CrossRef] [PubMed]
- Pelletier, B.; Trott, A.; Morano, K.A.; Labbé, S. Functional characterization of the iron-regulatory transcription factor Fep1 from Schizosaccharomyces pombe. J. Biol. Chem. 2005, 280, 25146–25161. [Google Scholar] [CrossRef]
- Ogunjimi, B.; Zhang, S.-Y.; Sørensen, K.B.; Skipper, K.A.; Carter-Timofte, M.; Kerner, G.; Luecke, S.; Prabakaran, T.; Cai, Y.; Meester, J. Inborn errors in RNA polymerase III underlie severe varicella zoster virus infections. J. Clin. Investig. 2017, 127, 3543–3556. [Google Scholar] [CrossRef]
- Girbig, M.; Misiaszek, A.D.; Vorländer, M.K.; Lafita, A.; Grötsch, H.; Baudin, F.; Bateman, A.; Müller, C.W. Cryo-EM structures of human RNA polymerase III in its unbound and transcribing states. Nat. Struct. Mol. Biol. 2021, 28, 210–219. [Google Scholar] [CrossRef]
- Wang, Q.; Li, S.; Wan, F.; Xu, Y.; Wu, Z.; Cao, M.; Lan, P.; Lei, M.; Wu, J. Structural insights into transcriptional regulation of human RNA polymerase III. Nat. Struct. Mol. Biol. 2021, 28, 220–227. [Google Scholar] [CrossRef]
- Wang, Z.; Roeder, R.G. Three human RNA polymerase III-specific subunits form a subcomplex with a selective function in specific transcription initiation. Genes Dev. 1997, 11, 1315–1326. [Google Scholar] [CrossRef]
- Chiu, Y.-H.; MacMillan, J.B.; Chen, Z.J. RNA polymerase III detects cytosolic DNA and induces type I interferons through the RIG-I pathway. Cell 2009, 138, 576–591. [Google Scholar] [CrossRef]
- Tian, K.; Wang, R.; Huang, J.; Wang, H.; Ji, X. Subcellular localization shapes the fate of RNA polymerase III. Cell Rep. 2023, 42, 112941. [Google Scholar] [CrossRef]
- Zurita, M.; Merino, C. The transcriptional complexity of the TFIIH complex. TRENDS Genet. 2003, 19, 578–584. [Google Scholar] [CrossRef] [PubMed]
- Giglia-Mari, G.; Coin, F.; Ranish, J.A.; Hoogstraten, D.; Theil, A.; Wijgers, N.; Jaspers, N.G.; Raams, A.; Argentini, M.; Van Der Spek, P. A new, tenth subunit of TFIIH is responsible for the DNA repair syndrome trichothiodystrophy group A. Nat. Genet. 2004, 36, 714–719. [Google Scholar] [CrossRef] [PubMed]
- Keriel, A.; Stary, A.; Sarasin, A.; Rochette-Egly, C.; Egly, J.-M. XPD mutations prevent TFIIH-dependent transactivation by nuclear receptors and phosphorylation of RARα. Cell 2002, 109, 125–135. [Google Scholar] [CrossRef] [PubMed]
- Dubaele, S.; De Santis, L.P.; Bienstock, R.J.; Keriel, A.; Stefanini, M.; Van Houten, B.; Egly, J.-M. Basal transcription defect discriminates between xeroderma pigmentosum and trichothiodystrophy in XPD patients. Mol. Cell 2003, 11, 1635–1646. [Google Scholar] [CrossRef] [PubMed]
- Compe, E.; Drané, P.; Laurent, C.; Diderich, K.; Braun, C.; Hoeijmakers, J.H.; Egly, J.-M. Dysregulation of the peroxisome proliferator-activated receptor target genes by XPD mutations. Mol. Cell. Biol. 2005, 25, 6065–6076. [Google Scholar] [CrossRef]
- Abbassi, N.-e.-H.; Jaciuk, M.; Scherf, D.; Böhnert, P.; Rau, A.; Hammermeister, A.; Rawski, M.; Indyka, P.; Wazny, G.; Chramiec-Głąbik, A. Cryo-EM structures of the human Elongator complex at work. Nat. Commun. 2024, 15, 4094. [Google Scholar] [CrossRef]
- Selvadurai, K.; Wang, P.; Seimetz, J.; Huang, R.H. Archaeal Elp3 catalyzes tRNA wobble uridine modification at C5 via a radical mechanism. Nat. Chem. Biol. 2014, 10, 810–812. [Google Scholar] [CrossRef]
- Dalwadi, U.; Yip, C.K. Structural insights into the function of Elongator. Cell. Mol. Life Sci. 2018, 75, 1613–1622. [Google Scholar] [CrossRef]
- Maio, N.; Orbach, R.; Zaharieva, I.T.; Töpf, A.; Donkervoort, S.; Munot, P.; Mueller, J.; Willis, T.; Verma, S.; Peric, S. CIAO1 loss of function causes a neuromuscular disorder with compromise of nucleocytoplasmic Fe-S enzymes. J. Clin. Investig. 2024, 134, e179559. [Google Scholar] [CrossRef]
- Romero, A.M.; Martinez-Pastor, M.T.; Puig, S. Iron in Translation: From the Beginning to the End. Microorganisms 2021, 9, 1058. [Google Scholar] [CrossRef]
- Sun, C.; Guo, R.; Ye, X.; Tang, S.; Chen, M.; Zhou, P.; Yang, M.; Liao, C.; Li, H.; Lin, B. Wybutosine hypomodification of tRNAphe activates HERVK and impairs neuronal differentiation. Iscience 2024, 27, 109748. [Google Scholar] [CrossRef] [PubMed]
- Kessler, A.C.; Silveira d’Almeida, G.; Alfonzo, J.D. The role of intracellular compartmentalization on tRNA processing and modification. RNA Biol. 2018, 15, 554–566. [Google Scholar] [CrossRef] [PubMed]
- Wei, F.Y.; Zhou, B.; Suzuki, T.; Miyata, K.; Ujihara, Y.; Horiguchi, H.; Takahashi, N.; Xie, P.; Michiue, H.; Fujimura, A.; et al. Cdk5rap1-mediated 2-methylthio modification of mitochondrial tRNAs governs protein translation and contributes to myopathy in mice and humans. Cell Metab. 2015, 21, 428–442. [Google Scholar] [CrossRef]
- Sheng, L.; Hao, S.-L.; Yang, W.-X.; Sun, Y. The multiple functions of kinesin-4 family motor protein KIF4 and its clinical potential. Gene 2018, 678, 90–99. [Google Scholar] [CrossRef]
- Lee, Y.M.; Kim, W. Kinesin superfamily protein member 4 (KIF4) is localized to midzone and midbody in dividing cells. Exp. Mol. Med. 2004, 36, 93–97. [Google Scholar] [CrossRef]
- Mazumdar, M.; Sundareshan, S.; Misteli, T. Human chromokinesin KIF4A functions in chromosome condensation and segregation. J. Cell Biol. 2004, 166, 613–620. [Google Scholar] [CrossRef]
- Zhu, C.; Jiang, W. Cell cycle-dependent translocation of PRC1 on the spindle by Kif4 is essential for midzone formation and cytokinesis. Proc. Natl. Acad. Sci. USA 2005, 102, 343–348. [Google Scholar] [CrossRef]
- Kurasawa, Y.; Earnshaw, W.C.; Mochizuki, Y.; Dohmae, N.; Todokoro, K. Essential roles of KIF4 and its binding partner PRC1 in organized central spindle midzone formation. EMBO J. 2004, 23, 3237–3248. [Google Scholar] [CrossRef]
- Mazumdar, M.; Sung, M.-H.; Misteli, T. Chromatin maintenance by a molecular motor protein. Nucleus 2011, 2, 591–600. [Google Scholar] [CrossRef]
- Wu, G.; Zhou, L.; Khidr, L.; Guo, X.E.; Kim, W.; Lee, Y.M.; Krasieva, T.; Chen, P.-L. A novel role of the chromokinesin Kif4A in DNA damage response. Cell Cycle 2008, 7, 2013–2020. [Google Scholar] [CrossRef]
- Kalantari, S.; Carlston, C.; Alsaleh, N.; Abdel-Salam, G.M.; Alkuraya, F.; Kato, M.; Matsumoto, N.; Miyatake, S.; Yamamoto, T.; Fares-Taie, L. Expanding the KIF4A-associated phenotype. Am. J. Med. Genet. Part A 2021, 185, 3728–3739. [Google Scholar] [CrossRef] [PubMed]
- Gad, S.A.; Sugiyama, M.; Tsuge, M.; Wakae, K.; Fukano, K.; Oshima, M.; Sureau, C.; Watanabe, N.; Kato, T.; Murayama, A. The kinesin KIF4 mediates HBV/HDV entry through the regulation of surface NTCP localization and can be targeted by RXR agonists in vitro. PLoS Pathog. 2022, 18, e1009983. [Google Scholar] [CrossRef] [PubMed]
- Mazumdar, M.; Misteli, T. Chromokinesins: Multitalented players in mitosis. Trends Cell Biol. 2005, 15, 349–355. [Google Scholar] [CrossRef]
- Wu, G.; Chen, P.-L. Structural requirements of chromokinesin Kif4A for its proper function in mitosis. Biochem. Biophys. Res. Commun. 2008, 372, 454–458. [Google Scholar] [CrossRef]
- Barton, J.K.; Silva, R.M.; O’Brien, E. Redox chemistry in the genome: Emergence of the [4Fe4S] cofactor in repair and replication. Annu. Rev. Biochem. 2019, 88, 163–190. [Google Scholar] [CrossRef]
- Ueda, C.; Langton, M.; Chen, J.; Pandelia, M.-E. The HBx protein from hepatitis B virus coordinates a redox-active Fe-S cluster. J. Biol. Chem. 2022, 298, 101698. [Google Scholar] [CrossRef]
- Henkler, F.; Hoare, J.; Waseem, N.; Goldin, R.D.; McGarvey, M.J.; Koshy, R.; King, I.A. Intracellular localization of the hepatitis B virus HBx protein. J. Gen. Virol. 2001, 82, 871–882. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2024 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Novoa-Aponte, L.; Leon-Torres, A.; Philpott, C.C. Guardians of the Genome: Iron–Sulfur Proteins in the Nucleus. Inorganics 2024, 12, 316. https://doi.org/10.3390/inorganics12120316
Novoa-Aponte L, Leon-Torres A, Philpott CC. Guardians of the Genome: Iron–Sulfur Proteins in the Nucleus. Inorganics. 2024; 12(12):316. https://doi.org/10.3390/inorganics12120316
Chicago/Turabian StyleNovoa-Aponte, Lorena, Andres Leon-Torres, and Caroline C. Philpott. 2024. "Guardians of the Genome: Iron–Sulfur Proteins in the Nucleus" Inorganics 12, no. 12: 316. https://doi.org/10.3390/inorganics12120316
APA StyleNovoa-Aponte, L., Leon-Torres, A., & Philpott, C. C. (2024). Guardians of the Genome: Iron–Sulfur Proteins in the Nucleus. Inorganics, 12(12), 316. https://doi.org/10.3390/inorganics12120316