Hydroxyurea-The Good, the Bad and the Ugly

Marcelina W Musiałek, Dorota Rybaczek, Marcelina W Musiałek, Dorota Rybaczek

Abstract

Hydroxyurea (HU) is mostly referred to as an inhibitor of ribonucleotide reductase (RNR) and as the agent that is commonly used to arrest cells in the S-phase of the cycle by inducing replication stress. It is a well-known and widely used drug, one which has proved to be effective in treating chronic myeloproliferative disorders and which is considered a staple agent in sickle anemia therapy and-recently-a promising factor in preventing cognitive decline in Alzheimer's disease. The reversibility of HU-induced replication inhibition also makes it a common laboratory ingredient used to synchronize cell cycles. On the other hand, prolonged treatment or higher dosage of hydroxyurea causes cell death due to accumulation of DNA damage and oxidative stress. Hydroxyurea treatments are also still far from perfect and it has been suggested that it facilitates skin cancer progression. Also, recent studies have shown that hydroxyurea may affect a larger number of enzymes due to its less specific interaction mechanism, which may contribute to further as-yet unspecified factors affecting cell response. In this review, we examine the actual state of knowledge about hydroxyurea and the mechanisms behind its cytotoxic effects. The practical applications of the recent findings may prove to enhance the already existing use of the drug in new and promising ways.

Keywords: DNA replication checkpoint; cell cycle arrest; hydroxyurea; replication stress; ribonucleotide reductase.

Conflict of interest statement

The authors declare no conflict of interest. The founder had no role in the design of the study; in the data collection, analysis and interpretation; in the writing of the manuscript, or in the decision to publish.

Figures

Figure 1
Figure 1
The structures of two homodimeric subunits of ribonucleotide reductase and the impact of hydroxyurea on the PCET transfer: (A) the dATP in the allosteric center of subunit α2 of human RNR (PDB: 2WGH) [36]; (B) the Fe(3+) ion center of the human β2 subunit (PDB: 2UW2); (C) a 3D model of the α2 subunit with the allosteric sites and active (substrate) sites (PDB: 2WGH); (D) a 3D model of the β2 subunit with the di-iron center placement marked (PDB: 2UW2). (E) Bidirectional PCET can occur only when both subunits form an α2β2 heterodimeric structure. This binding allows for substrate–effector interaction. (F) Hydroxyurea inhibits PCET within RNR by proton-coupled electron transfer, most probably mediated by a hydrogen-bonded proton wire [35]. The images in (A,B) were generated using Mol*Viewer [46].
Figure 2
Figure 2
The cell response to HU-induced depletion of dNTPs during replication. Origins licensed on the DNA strand are grouped into replication “factories” the activation of which is strictly defined in time, from early- to late-replicating. In normal conditions, only a small percentage of origins will become activated and start to replicate DNA; the other origins remain dormant and will be replicated passively unless they are needed (A). If the currently replicating factory already has the desired amount of active forks, an ATR/Chk1 signaling mechanism limits dormant origin firing and also limits the activation of other factories that are supposed to replicate later (C). In the absence of free dNTPs (B), active forks stall and expose fragile ssDNA (B,C). ATR kinases stabilize stalled forks and also trigger the signaling pathway of the S-phase checkpoint, which arrests the cell cycle, inhibits origin firing in inactive factories, and restores RNR activity (C). The cell is unable to enter mitosis unless DNA replication is finished and any damage repaired (C).
Figure 3
Figure 3
HU changes the activity of replication as well as several key proteins in root meristem cells of Vicia faba. (A) Nuclear DNA labeled with 5-bromo-2′-deoxyuridine (BrdU; a1a3); (B) PCNA (b1b3); (C) Chk1 kinase (c1c3). The DNA within the nucleus is detected with propidium iodide (orange; c1). The control levels of Chk1 are consistent with reports that cells maintain a basic level of kinases in order to prevent dormant origin firing (c1,c3). (D) Histone H2AX (d1d3), which is a hallmark of double-strand DNA damage (H2AXS139ph), shows increased levels after HU exposure (d2,d3). The presented factors were detected using immunocytochemical staining after 32 h incubation in water (first column, normal conditions; a1,b1,c1,d1) or in 2.5 mM HU (second column; a2,b2,c2,d2). The third column contains 3D renderings of BrdU/protein localization within the nucleus, based on statistical analysis of experimental data and microscope observations (a3,b3,c3,d3). The left side of the column shows the agent’s activity under normal conditions and the right panel shows the activity after HU treatment. The experimental procedures for immunocytochemical detection of: (i) PCNA, (ii) Chk1 (phosphorylated on serine 317), and (iii) H2AX (phosphorylated on serine 139) were identical (BrdU detection required hydrolysis with HCl). The experimental procedure for the DNA replication assay with BrdU (on entire cells) was as follows: control and HU-treated (2.5 mM, 24 h) seedlings were pulsed for 30 min with 30 µM BrdU solution at 20 °C in the dark. Excised 3 mm long meristems were then washed with ice-cold Tris buffer (10 mM Tris, 10 mM EDTA-2Na, 100 mM NaCl, pH 7.2) for 5 min and fixed for 45 min at 4 °C in freshly prepared 4% paraformaldehyde. After fixation, root tips were washed, squashed onto slides, and treated with 1.5 M HCl (for 1.5 h at 20 °C, for partial denaturation of nuclear DNA). In contrast, detection of PCNA/Chk1S317ph/H2AXS139ph required an enzymatic maceration step, i.e., incubation in a citric acid-buffered digestion solution for 45 min at 37 °C. Subsequently, root tips were washed again, squashed onto slides, and treated with the following primary antibodies: (i) mouse monoclonal anti-BrdU (Sigma-Aldrich, Saint Quentin, France) diluted in Tris-buffer (1:50) or (ii) the rabbit monoclonal antibodies anti-PCNA (Abcam, Cambridge, United Kingdom), anti-Chk1(S317ph) (Cell Signaling Technology, Beverly, MA, USA), and anti-H2AX(S139ph) (Cell Signaling Technology; Beverly, MA, USA; 1:250), diluted in PBS. Following overnight incubation (at least 16 h) at 4 °C, slides were washed in Tris/PBS buffer, respectively, and incubated for 1 h with AlexaFluor 488-conjugated goat anti-mouse/mouse anti-rabbit secondary antibodies, respectively (Cell Signaling Technology, Beverly, MA, USA; 1:500), washed, and embedded in PBS/glycerol mixture (9:1) with 3% DABCO.
Figure 4
Figure 4
The consequences of ATR inhibition and malfunction of the S-phase checkpoint. A disrupted ATR/Chk1 pathway does not arrest the cell in the S phase and allows cycle progression with under-replicated DNA. This causes the accumulation of SSBs and DSBs, fork collapse, and deregulation of the inhibition of late origin firing. Moreover, a lack of radical maintenance and inhibition of new RNR synthesis leads to the accumulation of HU-induced ROS. The circled images show anaphase aberrations detected using Feulgen staining (top) and the accumulation of ROS (H2O2) in the cytoplasm in DAB-stained root meristem cells [99] (bottom) of Vicia faba subsp. minor. The experimental procedure for 3,3-diaminobenzidine (DAB) staining (to detect H2O2 by means of DAB polymerization) was as follows: roots of V. faba were submerged in Tris-buffered (10 mM Tris, 10 mM EDTA-2Na, 100 mM NaCl) DAB-HCl (1 mg mL−1; pH 7.5) dissolved in distilled water or in 2.5 mM HU (HU-treated plants). Following 24 h incubation, excised 3 mm root tips were fixed for 45 min in PBS-buffered 4% paraformaldehyde (20 °C), washed (3 times) with PBS, and placed in citric acid-buffered 2.5% pectinase (pH 5.0; 37 °C for 45 min). Digested root meristem cells were washed with PBS and squashed onto microscope slides in a mixture of glycerol and PBS (9:1; v/v). Production of H2O2 was observed under the microscope as reddish-brown areas in the cells (indicated by a black arrow in the circled image).
Figure 5
Figure 5
Visual examples of cell death, DNA damage, and chromosome breakage induced by overexposure to HU or bypassing of the S-phase cycle induced by 2.5 mM HU and 5 mM caffeine (CF) in root meristem cells of Vicia faba subsp. minor. (A) Programmed cell death (PCD) visible after double staining with acridine orange (AO) and ethidium bromide (EB) (a1,a2) [100]. AO is able to enter living cells and makes them appear green (a1). EB is taken up only when cell membrane integrity is compromised (usually in dead or dying cells), making them look red (a2). The range of colors (from green through yellow to red) indicates the existence of living, dying, and dead cells. The scale bar equals 50 µm. (B) DNA damage observed under normal conditions (b1) and after HU treatment (b2). The longer and more visible the comet’s tail is, the more DNA damage is present. DNA was stained with YOYO-1 (green; b1b3). The scale bar equals 10 µm. (C) Feulgen-stained DNA and chromosome aberrations during mitosis, visible under a red spectrum in a fluorescent microscope (c1c4) [101]. The experimental procedure for Feulgen staining (d1d4,e1,e2) was as follows: root tips were fixed in cold absolute ethanol and glacial acetic acid (3:1, v/v) for 1 h, washed several times with ethanol, rehydrated, hydrolyzed in 4 M HCl (1.5 h), and stained with Schiff’s reagent (pararosaniline; Sigma-Aldrich, Saint Quentin, France) according to standard methods. After rising in SO2–water (three times) and distilled water, 1.5 mm long apical segments of roots were cut off, placed in a drop of 45% acetic acid, and squashed onto microscope slides. Following freezing using dry ice, coverslips were removed, and the dehydrated slides were mounted in Canada balsam. The top right inserts show mitotic morphology under normal conditions; DNA was also Feulgen-stained and the fluorescent image colors were edited to green for better visibility. The scale bars equal 10 µm. (D) Additional examples of anaphase aberrations (Feulgen staining under visible light). The single black asterisks (*) indicate mitotic chromosome fragments that were lost due to a lack of a connection to kinetochores (d1,d3); the double black asterisks (**) indicate O-shaped chromosomes (d2); the three asterisks (***) indicate a chromosome bridge (d4). The scale bars equal 10 µm. (E) Post-mitotic aberrations of decondensing DNA (Feulgen staining under visible light). Chromosome fragments (lost between opposite poles of the cell), i.e., later micronuclei, are indicated with a red asterisk (e1,e2). The scale bars equal 10 µm. (F) Short schematic of an S-phase checkpoint malfunction caused by caffeine (CF), as used in the presented experiments.
Figure 6
Figure 6
A summary of (A) the effects of HU exposure; (B) the consequences of high dosage, long treatment, and checkpoint deregulation; and (C) the medical and scientific applications of hydroxyurea.

References

    1. Rosenthal F., Wislicki L., Kollek L. Über die Beziehungen von Schwersten Blutgiften zu Abbauprodukten des Eiweiss—Ein Beitrag zum Entstehungsmechanismus der perniziösen Anämie. Klin. Wochenschr. 1928;7:972–977. doi: 10.1007/BF01716922.
    1. Adamson R.H. Activity of Congeners of Hydroxyurea Against Advanced Leukemia L1210. Proc. Soc. Exp. Biol. Med. 1965;119:456–458. doi: 10.3181/00379727-119-30209.
    1. Stearns B., Losee K.A., Bernstein J. Hydroxyurea: A New Type of Potential Antitumor Agent. J. Med. Chem. 1963;6:201. doi: 10.1021/jm00338a026.
    1. Madaan K., Kaushik D., Verma T. Hydroxyurea: A key player in cancer chemotherapy. Expert Rev. Anticancer Ther. 2012;12:19–29. doi: 10.1586/era.11.175.
    1. Spivak J.L., Hasselbalch H. Hydroxycarbamide: A user’s guide for chronic myeloproliferative disorders. Expert Rev. Anticancer Ther. 2011;11:403–414. doi: 10.1586/era.11.10.
    1. Yogev O., Anzi S., Inoue K., Shaulian E. Induction of transcriptionally active Jun proteins regulates drug-induced senescence. J. Biol. Chem. 2006;281:34475–34483. doi: 10.1074/jbc.M602865200.
    1. Nevitt S.J., Jones A.P., Howard J. Hydroxyurea (hydroxycarbamide) for sickle cell disease. Cochrane Database Syst. Rev. 2017;4:CD002202. doi: 10.1002/14651858.CD002202.pub2.
    1. Tshilolo L., Tomlinson G., Williams T.N., Santos B., Olupot-Olupot P., Lane A., Aygun B., Stuber S.E., Latham T.S., McGann P.T., et al. Hydroxyurea for Children with Sickle Cell Anemia in Sub-Saharan Africa. N. Engl. J. Med. 2019;380:121–131. doi: 10.1056/NEJMoa1813598.
    1. Brose R.D., Lehrmann E., Zhang Y., Reeves R.H., Smith K.D., Mattson M.P. Hydroxyurea attenuates oxidative, metabolic, and excitotoxic stress in rat hippocampal neurons and improves spatial memory in a mouse model of Alzheimer’s disease. Neurobiol. Aging. 2018;72:121–133. doi: 10.1016/j.neurobiolaging.2018.08.021.
    1. Koç A., Wheeler L.J., Mathews C.K., Merrill G.F. Hydroxyurea Arrests DNA Replication by a Mechanism that Preserves Basal dNTP Pools. J. Biol. Chem. 2004;279:223–230. doi: 10.1074/jbc.M303952200.
    1. Berniak K., Rybak P., Bernas T., Zarebski M., Biela E., Zhao H., Darzynkiewicz Z., Dobrucki J.W. Relationship between DNA damage response, initiated by camptothecin or oxidative stress, and DNA replication, analyzed by quantitative 3D image analysis. Cytom. Part. A. 2013;83:913–924. doi: 10.1002/cyto.a.22327.
    1. Sarkaria J.N., Busby E.C., Tibbetts R.S., Roos P., Taya Y., Karnitz L.M., Abraham R.T. Inhibition of ATM and ATR Kinase Activities by the Radiosensitizing Agent, Caffeine. Cancer Res. 1999;59:4375–4382.
    1. Boddy M.N., Russell P. DNA replication checkpoint. Curr. Biol. 2001;11:R953–R956. doi: 10.1016/S0960-9822(01)00572-3.
    1. Rybaczek D. Ultrastructural changes associated with the induction of premature chromosome condensation in Vicia faba root meristem cells. Plant. Cell Rep. 2014;33:1547–1564. doi: 10.1007/s00299-014-1637-0.
    1. Xu Y.J., Singh A., Alter G.M. Hydroxyurea induces cytokinesis arrest in cells expressing a mutated sterol-14α-demethylase in the ergosterol biosynthesis pathway. Genetics. 2016;204:959–973. doi: 10.1534/genetics.116.191536.
    1. Huang M.E., Facca C., Fatmi Z., Baïlle D., Bénakli S., Vernis L. DNA replication inhibitor hydroxyurea alters Fe-S centers by producing reactive oxygen species in vivo. Sci. Rep. 2016;6:29361. doi: 10.1038/srep29361.
    1. King S.B. Nitric oxide production from hydroxyurea. Free Radic. Biol. Med. 2004;37:737–744. doi: 10.1016/j.freeradbiomed.2004.02.073.
    1. Ciccia A., Elledge S.J. The DNA Damage Response: Making It Safe to Play with Knives. Mol. Cell. 2010;40:179–204. doi: 10.1016/j.molcel.2010.09.019.
    1. Trovesi C., Manfrini N., Falcettoni M., Longhese M.P. Regulation of the DNA damage response by cyclin-dependent kinases. J. Mol. Biol. 2013;425:4756–4766. doi: 10.1016/j.jmb.2013.04.013.
    1. Huh M.S., Ivanochko D., Hashem L.E., Curtin M., Delorme M., Goodall E., Yan K., Picketts D.J. Stalled replication forks within heterochromatin require ATRX for protection. Cell Death Dis. 2016;7:e2220-12. doi: 10.1038/cddis.2016.121.
    1. Giannattasio M., Branzei D. S-phase checkpoint regulations that preserve replication and chromosome integrity upon dNTP depletion. Cell. Mol. Life Sci. 2017;74:2361–2380. doi: 10.1007/s00018-017-2474-4.
    1. Moiseeva T.N., Yin Y., Calderon M.J., Qian C., Schamus-Haynes S., Sugitani N., Osmanbeyoglu H.U., Rothenberg E., Watkins S.C., Bakkenist C.J. An ATR and CHK1 kinase signaling mechanism that limits origin firing during unperturbed DNA replication. Proc. Natl. Acad. Sci. USA. 2019;116:13374–13383. doi: 10.1073/pnas.1903418116.
    1. Nair J., Huang T.T., Murai J., Haynes B., Steeg P.S., Pommier Y., Lee J.M. Resistance to the CHK1 inhibitor prexasertib involves functionally distinct CHK1 activities in BRCA wild-type ovarian cancer. Oncogene. 2020;39:5520–5535. doi: 10.1038/s41388-020-1383-4.
    1. Julius J., Peng J., McCulley A., Caridi C., Arnak R., See C., Nugent C.I., Feng W., Bachant J. Inhibition of spindle extension through the yeast S phase checkpoint is coupled to replication fork stability and the integrity of centromeric DNA. Mol. Biol. Cell. 2019;30:2771–2789. doi: 10.1091/mbc.E19-03-0156.
    1. Shao J., Zhou B., Chu B., Yen Y. Ribonucleotide Reductase Inhibitors and Future Drug Design. Curr. Cancer Drug Targets. 2006;6:409–431. doi: 10.2174/156800906777723949.
    1. Sethy S., Panda T., Jena R.K. Beneficial Effect of Low Fixed Dose of Hydroxyurea in Vaso-occlusive Crisis and Transfusion Requirements in Adult HbSS Patients: A Prospective Study in a Tertiary Care Center. Indian J. Hematol. Blood Transfus. 2018;34:294–298. doi: 10.1007/s12288-017-0869-x.
    1. Yazinski S.A., Zou L. Functions, Regulation, and Therapeutic Implications of the ATR Checkpoint Pathway. Annu. Rev. Genet. 2016;50:155–173. doi: 10.1146/annurev-genet-121415-121658.
    1. Ashley A.K., Shrivastav M., Nie J., Amerin C., Troksa K., Glanzer J.G., Liu S., Opiyo S.O., Dimitrova D.D., Le P., et al. DNA-PK phosphorylation of RPA32 Ser4/Ser8 regulates replication stress checkpoint activation, fork restart, homologous recombination and mitotic catastrophe. DNA Repair. 2014;21:131–139. doi: 10.1016/j.dnarep.2014.04.008.
    1. Kramara J., Osia B., Malkova A. Break Induced Replication: The where, the why, and the how. HHS Public Access. Trends Genet. 2018;34:518–531. doi: 10.1016/j.tig.2018.04.002.
    1. Bester A.C., Roniger M., Oren Y.S., Im M.M., Sarni D., Chaoat M., Bensimon A., Zamir G., Shewach D.S., Kerem B. Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell. 2011;145:435–446. doi: 10.1016/j.cell.2011.03.044.
    1. Przybyszewski W.M., Kasperczyk J. Rodnikowy mechanizm ubocznej toksyczności hydroksymocznika. Postepy Hig. Med. Dosw. 2006;60:516–526.
    1. De Oliveira E.A.M., Boy K.d.A., Santos A.P.P., Machado C.d.S., Velloso-Rodrigues C., Gerheim P.S.A.S., Mendonça L.M. Evaluation of hydroxyurea genotoxicity in patients with sickle cell disease. Einstein. 2019;17:eAO4742. doi: 10.31744/einstein_journal/2019AO4742.
    1. Al Mamun M., Albergante L., Moreno A., Carrington J.T., Blow J.J., Newman T.J. Inevitability and containment of replication errors for eukaryotic genome lengths spanning megabase to gigabase. Proc. Natl. Acad. Sci. USA. 2016;113:E5765–E5774. doi: 10.1073/pnas.1603241113.
    1. Elledge S.J., Zhou Z., Allen J.B. Ribonucleotide reductase: Regulation, regulation, regulation. Trends Biochem. Sci. 1992;17:119–123. doi: 10.1016/0968-0004(92)90249-9.
    1. Offenbacher A.R., Barry B.A. A Proton Wire Mediates Proton Coupled Electron Transfer from Hydroxyurea and Other Hydroxamic Acids to Tyrosyl Radical in Class Ia Ribonucleotide Reductase. J. Phys. Chem. B. 2020;124:345–354. doi: 10.1021/acs.jpcb.9b08587.
    1. Fairman J.W., Wijerathna S.R., Ahmad M.F., Xu H., Nakano R., Jha S., Prendergast J., Welin R.M., Flodin S., Roos A., et al. Structural basis for allosteric regulation of human ribonucleotide reductase by nucleotide-induced oligomerization. Nat. Struct. Mol. Biol. 2011;18:316–322. doi: 10.1038/nsmb.2007.
    1. Brignole E.J., Tsai K.L., Chittuluru J., Li H., Aye Y., Penczek P.A., Stubbe J.A., Drennan C.L., Asturias F. 3.3-Å resolution cryo-EM structure of human ribonucleotide reductase with substrate and allosteric regulators bound. eLife. 2018;7:e31502. doi: 10.7554/eLife.31502.
    1. Eriksson M., Uhlin U., Ramaswamy S., Ekberg M., Regnström K., Sjöberg B.M., Eklund H. Binding of allosteric effectors to ribonucleotide reductase protein R1: Reduction of active-site cysteines promotes substrate binding. Structure. 1997;5:1077–1092. doi: 10.1016/S0969-2126(97)00259-1.
    1. Rofougaran R., Vodnala M., Hofer A. Enzymatically active mammalian ribonucleotide reductase exists primarily as an α6β2 octamer. J. Biol. Chem. 2006;281:27705–27711. doi: 10.1074/jbc.M605573200.
    1. Radivoyevitch T. Automated mass action model space generation and analysis methods for two-reactant combinatorially complex equilibriums: An analysis of ATP-induced ribonucleotide reductase R1 hexamerization data. Biol. Direct. 2009;4:50. doi: 10.1186/1745-6150-4-50.
    1. Denysenkov V.P., Biglino D., Lubitz W., Prisner T.F., Bennati M. Structure of the tyrosyl biradical in mouse R2 ribonucleotide reductase from high-field PELDOR. Angew. Chem. Int. Ed. 2008;47:1224–1227. doi: 10.1002/anie.200703753.
    1. Minnihan E.C., Nocera D.G., Stubbe J. Reversible, long-range radical transfer in E. coli class Ia ribonucleotide reductase. Acc. Chem. Res. 2013;46:2524–2535. doi: 10.1021/ar4000407.
    1. Martínez-Carranza M., Jonna V.R., Lundin D., Sahlin M., Carlson L.A., Jemal N., Högbom M., Sjöberg B.M., Stenmark P., Hofer A. A ribonucleotide reductase from clostridium botulinum reveals distinct evolutionary pathways to regulation via the overall activity site. J. Biol. Chem. 2020;295:15576–15587. doi: 10.1074/jbc.RA120.014895.
    1. Kang G., Taguchi A.T., Stubbe J.A., Drennan C.L. Structure of a trapped radical transfer pathway within a ribonucleotide reductase holocomplex. Science. 2020;368:424–427. doi: 10.1126/science.aba6794.
    1. Nordlund P., Reichard P. Ribonucleotide reductases. Annu. Rev. Biochem. 2006;75:681–706. doi: 10.1146/annurev.biochem.75.103004.142443.
    1. Sehnal D., Bittrich S., Deshpande M., Svobodová R., Berka K., Bazgier V., Velenkar S., Burley S.K., Koča J., Rose A.S. Mol*Viewer: Modern web app for 3D visualization and analysis of large biomolecular structures. Nucleic Acids Res. 2021;49:W431–W437. doi: 10.1093/nar/gkab314.
    1. Rofougaran R., Crona M., Vodnala M., Sjöberg B.-M., Hofer A. Oligomerization status directs overall activity regulation of the Escherichia coli class Ia ribonucleotide reductase. J. Biol. Chem. 2008;283:35310–35318. doi: 10.1074/jbc.M806738200.
    1. Hofer A., Crona M., Logan D.T., Sjöberg B.-M. DNA building blocks: Keeping control of manufacture. Crit. Rev. Biochem. Mol. Biol. 2012;47:50–63. doi: 10.3109/10409238.2011.630372.
    1. Håkansson P., Hofer A., Thelander L. Regulation of mammalian ribonucleotide reduction and dNTP pools after DNA damage and in resting cells. J. Biol. Chem. 2006;281:7834–7841. doi: 10.1074/jbc.M512894200.
    1. Sanvisens N., De Llanos R., Puig S. Function and regulation of yeast ribonucleotide reductase: Cell cycle, genotoxic stress, and iron bioavailability. Biomed. J. 2013;36:51–58.
    1. Guarino E., Salguero I., Kearsey S.E. Cellular regulation of ribonucleotide reductase in eukaryotes. Semin. Cell Dev. Biol. 2014;30:97–103. doi: 10.1016/j.semcdb.2014.03.030.
    1. Lassmann G., Thelander L., Gräslund A. EPR stopped-flow studies of the reaction of the tyrosyl radical of protein R2 from ribonucleotide reductase with hydroxyurea. Biochem. Biophys. Res. Commun. 1992;188:879–887. doi: 10.1016/0006-291X(92)91138-G.
    1. Singh A., Xu Y.J. The cell killing mechanisms of hydroxyurea. Genes. 2016;7:99. doi: 10.3390/genes7110099.
    1. Fontecave M., Lepoivre M., Elleingand E., Gerez C., Guittet O. Resveratrol, a remarkable inhibitor of ribonucleotide reductase. FEBS Lett. 1998;421:277–279. doi: 10.1016/S0014-5793(97)01572-X.
    1. Rawson J.M.O., Roth M.E., Xie J., Daly M.B., Clouser C.L., Landman S.R., Reilly C.S., Bonnac L., Kim B., Patterson S.E., et al. Synergistic reduction of HIV-1 infectivity by 5-azacytidine and inhibitors of ribonucleotide reductase. Bioorganic Med. Chem. 2016;24:2410–2422. doi: 10.1016/j.bmc.2016.03.052.
    1. Li Z., Chen Q.Q., Lam C.W.K., Guo J.R., Zhang W.J., Wang C.Y., Wong V.K.W., Yao M.C., Zhang W. Investigation into perturbed nucleoside metabolism and cell cycle for elucidating the cytotoxicity effect of resveratrol on human lung adenocarcinoma epithelial cells. Chin. J. Nat. Med. 2019;17:608–615. doi: 10.1016/S1875-5364(19)30063-9.
    1. Neuhard J. Studies on the acid-soluble nucleotide pool in Escherichia coli. IV. Effects of hydroxyurea. BBA Sect. Nucleic Acids Protein Synth. 1967;145:1–6. doi: 10.1016/0005-2787(67)90647-8.
    1. Morganroth P.A., Hanawalt P.C. Role of DNA replication and repair in thymineless death in Escherichia coli. J. Bacteriol. 2006;188:5286–5288. doi: 10.1128/JB.00543-06.
    1. Kuong K.J., Kuzminov A. Cyanide, Peroxide and Nitric Oxide Formation in Solutions of Hydroxyurea Causes Cellular Toxicity and May Contribute to Its Therapeutic Potency. J. Mol. Biol. 2009;390:845–862. doi: 10.1016/j.jmb.2009.05.038.
    1. Fraser D.I., Liu K.T., Reid B.J., Hawkins E., Sevier A., Pyle M., Robinson J.W., Ouellette P.H.R., Ballantyne J.S. Widespread natural occurrence of hydroxyurea in animals. PLoS ONE. 2015;10:e0142890. doi: 10.1371/journal.pone.0142890.
    1. Hallmark L., Almeida L.E.F., Kamimura S., Smith M., Quezado Z.M.N. Nitric oxide and sickle cell disease—Is there a painful connection? Exp. Biol. Med. 2021;246:332–341. doi: 10.1177/1535370220976397.
    1. Grant G.D., Cook J.G. The Temporal Regulation of S Phase Proteins during G1. Adv. Exp. Med. Biol. 2017;1042:335–369. doi: 10.1007/978-981-10-6955-0_16.
    1. Kaykov A., Nurse P. The spatial and temporal organization of origin firing during the S-phase of fission yeast. Genome Res. 2015;25:391–401. doi: 10.1101/gr.180372.114.
    1. Shoaib M., Walter D., Gillespie P.J., Izard F., Fahrenkrog B., Lleres D., Lerdrup M., Johansen J.V., Hansen K., Julien E., et al. Histone H4K20 methylation mediated chromatin compaction threshold ensures genome integrity by limiting DNA replication licensing. Nat. Commun. 2018;9:3704. doi: 10.1038/s41467-018-06066-8.
    1. Blow J.J. Defects in the origin licensing checkpoint stresses cells exiting G0. J. Cell Biol. 2019;218:2080–2081. doi: 10.1083/jcb.201905181.
    1. McIntosh D., Blow J.J. Dormant origins, the licensing checkpoint, and the response to replicative stresses. Cold Spring Harb. Perspect. Biol. 2012;4:a012955. doi: 10.1101/cshperspect.a012955.
    1. Musiałek M.W., Rybaczek D. Behavior of replication origins in Eukaryota—Spatio-temporal dynamics of licensing and firing. Cell Cycle. 2015;14:2251–2264. doi: 10.1080/15384101.2015.1056421.
    1. Halliwell J.A., Gravells P., Bryant H.E. DNA Fiber Assay for the Analysis of DNA Replication Progression in Human Pluripotent Stem Cells. Curr. Protoc. Stem Cell Biol. 2020;54:e115. doi: 10.1002/cpsc.115.
    1. Liu Y., Wang L., Xu X., Yuan Y., Zhang B., Li Z., Xie Y., Yan R., Zheng Z., Ji J., et al. The intra-S phase checkpoint directly regulates replication elongation to preserve the integrity of stalled replisomes. Proc. Natl. Acad. Sci USA. 2021;118:e2019183118. doi: 10.1073/pnas.2019183118.
    1. Nitani N., Nakamura K., Nakagawa C., Masukata H., Nakagawa T. Regulation of DNA replication machinery by Mrc1 in fission yeast. Genetics. 2006;174:155–165. doi: 10.1534/genetics.106.060053.
    1. Mukherjee C., Tripathi V., Manolika E.M., Heijink A.M., Ricci G., Merzouk S., de Boer H.R., Demmers J., van Vugt M.A.T.M., Ray Chaudhuri A. RIF1 promotes replication fork protection and efficient restart to maintain genome stability. Nat. Commun. 2019;10:3287. doi: 10.1038/s41467-019-11246-1.
    1. Jasencakova Z., Scharf A.N.D., Ask K., Corpet A., Imhof A., Almouzni G., Groth A. Replication Stress Interferes with Histone Recycling and Predeposition Marking of New Histones. Mol. Cell. 2010;37:736–743. doi: 10.1016/j.molcel.2010.01.033.
    1. Nazaretyan S.A., Savic N., Sadek M., Hackert B.J., Courcelle J., Courcelle C.T. Replication rapidly recovers and continues in the presence of hydroxyurea in Escherichia coli. J. Bacteriol. 2018;200:e00713-17. doi: 10.1128/JB.00713-17.
    1. Ivessa A.S., Lenzmeier B.A., Bessler J.B., Goudsouzian L.K., Schnakenberg S.L., Zakian V.A. The Saccharomyces cerevisiae helicase Rrm3p facilitates replication past nonhistone protein-DNA complexes. Mol. Cell. 2003;12:1525–1536. doi: 10.1016/S1097-2765(03)00456-8.
    1. Ercilla A., Feu S., Aranda S., Llopis A., Brynjólfsdóttir S.H., Sørensen C.S., Toledo L.I., Agell N. Acute hydroxyurea-induced replication blockade results in replisome components disengagement from nascent DNA without causing fork collapse. Cell. Mol. Life Sci. 2019;77:735–749. doi: 10.1007/s00018-019-03206-1.
    1. Gardner N.J., Gillespie P.J., Carrington J.T., Shanks E.J., McElroy S.P., Haagensen E.J., Frearson J.A., Woodland A., Blow J.J. The High-Affinity Interaction between ORC and DNA that Is Required for Replication Licensing Is Inhibited by 2-Arylquinolin-4-Amines. Cell Chem. Biol. 2017;24:981–992.e4. doi: 10.1016/j.chembiol.2017.06.019.
    1. Heidinger-Pauli J.M., Ünal E., Guacci V., Koshland D. The Kleisin Subunit of Cohesin Dictates Damage-Induced Cohesion. Mol. Cell. 2008;31:47–56. doi: 10.1016/j.molcel.2008.06.005.
    1. Liu Q., Guntuku S., Cui X.S., Matsuoka S., Cortez D., Tamai K., Luo G., Carattini-Rivera S., DeMayo F., Bradley A., et al. Chk1 is an essential kinase that is regulated by Atr and required for the G2/M DNA damage checkpoint. Genes Dev. 2000;14:1448–1459. doi: 10.1101/gad.14.12.1448.
    1. Menolfi D., Delamarre A., Lengronne A., Pasero P., Branzei D. Essential Roles of the Smc5/6 Complex in Replication through Natural Pausing Sites and Endogenous DNA Damage Tolerance. Mol. Cell. 2015;60:835–846. doi: 10.1016/j.molcel.2015.10.023.
    1. Weinert T.A., Kiser G.L., Hartwell L.H. Mitotic checkpoint genes in budding yeast and the dependence of mitosis on DNA replication and repair. Genes Dev. 1994;8:652–665. doi: 10.1101/gad.8.6.652.
    1. Zuilkoski C.M., Skibbens R.V. PCNA antagonizes cohesin-dependent roles in genomic stability. PLoS ONE. 2020;15:e0235103. doi: 10.1371/journal.pone.0235103.
    1. Van C., Yan S., Michael W.M., Waga S., Cimprich K.A. Continued primer synthesis at stalled replication forks contributes to checkpoint activation. J. Cell Biol. 2010;189:233–246. doi: 10.1083/jcb.200909105.
    1. Byun T.S., Pacek M., Yee M.C., Walter J.C., Cimprich K.A. Functional uncoupling of MCM helicase and DNA polymerase activities activates the ATR-dependent checkpoint. Genes Dev. 2005;19:1040–1052. doi: 10.1101/gad.1301205.
    1. Iyer D.R., Rhind N. Replication Fork Slowing and Stalling are Distinct, Checkpoint-Independent Consequences of Replicating Damaged DNA. PLoS Genet. 2017;13:e1006958. doi: 10.1371/journal.pgen.1006958.
    1. Zhao X., Chabes A., Domkin V., Thelander L., Rothstein R. The ribonucleotide reductase inhibitor Sml1 is a new target of the Mec1/Rad53 kinase cascade during growth and in response to DNA damage. EMBO J. 2001;20:3544–3553. doi: 10.1093/emboj/20.13.3544.
    1. Zhao X., Muller E.G.D., Rothstein R. A suppressor of two essential checkpoint genes identifies a novel protein that negatively affects dNTP pools. Mol. Cell. 1998;2:329–340. doi: 10.1016/S1097-2765(00)80277-4.
    1. Zhang Z., Reese J.C. Molecular Genetic Analysis of the Yeast Repressor Rfx1/Crt1 Reveals a Novel Two-Step Regulatory Mechanism. Mol. Cell. Biol. 2005;25:7399–7411. doi: 10.1128/MCB.25.17.7399-7411.2005.
    1. Woolstencroft R.N., Bellharz T.H., Cook M.A., Preiss T., Durocher D., Tyers M. Ccr4 contributes to tolerance of replication stress through control of CRT1 mRNA poly(A) tail length. J. Cell Sci. 2006;119:5178–5192. doi: 10.1242/jcs.03221.
    1. Buisson R., Boisvert J.L., Benes C.H., Zou L. Distinct but Concerted Roles of ATR, DNA-PK, and Chk1 in Countering Replication Stress during S Phase. Mol. Cell. 2015;59:1011–1024. doi: 10.1016/j.molcel.2015.07.029.
    1. Koppenhafer S.L., Goss K.L., Terry W.W., Gordon D.J. Inhibition of the ATR-CHK1 pathway in ewing sarcoma cells causes DNA damage and apoptosis via the CDK2-mediated degradation of RRM2. Mol. Cancer Res. 2020;18:91–104. doi: 10.1158/1541-7786.MCR-19-0585.
    1. Lopez-Contreras A.J., Specks J., Barlow J.H., Ambrogio C., Desler C., Vikingsson S., Rodrigo-Perez S., Green H., Rasmussen L.J., Murga M., et al. Increased Rrm2 gene dosage reduces fragile site breakage and prolongs survival of ATR mutant mice. Genes Dev. 2015;29:690–695. doi: 10.1101/gad.256958.114.
    1. Zou L., Elledge S.J. Sensing DNA Damage Through ATRIP Recognition of RPA-ssDNA Complexes. Science. 2003;300:1542–1548. doi: 10.1126/science.1083430.
    1. Frattini C., Villa-Hernández S., Pellicanò G., Jossen R., Katou Y., Shirahige K., Bermejo R. Cohesin Ubiquitylation and Mobilization Facilitate Stalled Replication Fork Dynamics. Mol. Cell. 2017;68:758–772.e4. doi: 10.1016/j.molcel.2017.10.012.
    1. Litwin I., Pilarczyk E., Wysocki R. The emerging role of cohesin in the DNA damage response. Genes. 2018;9:581. doi: 10.3390/genes9120581.
    1. Rhodes J.D.P., Haarhuis J.H.I., Grimm J.B., Rowland B.D., Lavis L.D., Nasmyth K.A. Cohesin Can Remain Associated with Chromosomes during DNA Replication. Cell Rep. 2017;20:2749–2755. doi: 10.1016/j.celrep.2017.08.092.
    1. Bot C., Pfeiffer A., Giordano F., Manjeera D.E., Dantuma N.P., Ström L. Independent mechanisms recruit the cohesin loader protein NIPBL to sites of DNA damage. J. Cell Sci. 2017;130:1134–1146. doi: 10.1242/jcs.197236.
    1. Toledo L.I., Altmeyer M., Rask M.B., Lukas C., Larsen D.H., Povlsen L.K., Bekker-Jensen S., Mailand N., Bartek J., Lukas J. XATR prohibits replication catastrophe by preventing global exhaustion of RPA. Cell. 2013;155:1088. doi: 10.1016/j.cell.2013.10.043.
    1. Hoffman E.A., McCulley A., Haarer B., Arnak R., Feng W. Break-seq reveals hydroxyurea-induced chromosome fragility as a result of unscheduled conflict between DNA replication and transcription. Genome Res. 2015;25:402–412. doi: 10.1101/gr.180497.114.
    1. Żabka A., Polit J.T., Maszewski J. DNA replication stress induces deregulation of the cell cycle events in root meristems of Allium cepa. Ann. Bot. 2012;110:1581–1591. doi: 10.1093/aob/mcs215.
    1. Rybaczek D., Musialek M.W., Balcerczyk A. Caffeine-induced premature chromosome condensation results in the apoptosis-like programmed cell death in root meristems of Vicia faba. PLoS ONE. 2015;10:e0142307. doi: 10.1371/journal.pone.0142307.
    1. Rybaczek D., Musiałek M.W., Vrána J., Petrovská B., Pikus E.G., Doležel J. Kinetics of DNA Repair in Vicia faba Meristem Regeneration Following Replication Stress. Cells. 2021;10:88. doi: 10.3390/cells10010088.
    1. Palumbo E., Matricardi L., Tosoni E., Bensimon A., Russo A. Replication dynamics at common fragile site FRA6E. Chromosoma. 2010;119:575–587. doi: 10.1007/s00412-010-0279-4.
    1. Hashash N., Johnson A.L., Cha R.S. Regulation of fragile sites expression in budding yeast by MEC1, RRM3 and hydroxyurea. J. Cell Sci. 2011;124:181–185. doi: 10.1242/jcs.077313.
    1. Técher H., Koundrioukoff S., Carignon S., Wilhelm T., Millot G.A., Lopez B.S., Brison O., Debatisse M. Signaling from Mus81-Eme2-Dependent DNA Damage Elicited by Chk1 Deficiency Modulates Replication Fork Speed and Origin Usage. Cell Rep. 2016;14:1114–1127. doi: 10.1016/j.celrep.2015.12.093.
    1. Cortez D. Replication-Coupled DNA Repair. Mol. Cell. 2019;74:866–876. doi: 10.1016/j.molcel.2019.04.027.
    1. Shiu J.L., Wu C.K., Chang S.B., Sun Y.J., Chen Y.J., Lai C.C., Chiu W.T., Chang W.T., Myung K., Su W.P., et al. The HLTF–PARP1 interaction in the progression and stability of damaged replication forks caused by methyl methanesulfonate. Oncogenesis. 2020;9:104. doi: 10.1038/s41389-020-00289-5.
    1. Couch F.B., Cortez D. Fork reversal, too much of a good thing. Cell Cycle. 2014;13:1049–1050. doi: 10.4161/cc.28212.
    1. Barlow J.H., Rothstein R. Rad52 recruitment is DNA replication independent and regulated by Cdc28 and the Mec1 kinase. EMBO J. 2009;28:1121–1130. doi: 10.1038/emboj.2009.43.
    1. Fumasoni M., Zwicky K., Vanoli F., Lopes M., Branzei D. Error-Free DNA Damage Tolerance and Sister Chromatid Proximity during DNA Replication Rely on the Polα/Primase/Ctf4 Complex. Mol. Cell. 2015;57:812–823. doi: 10.1016/j.molcel.2014.12.038.
    1. Zellweger R., Dalcher D., Mutreja K., Berti M., Schmid J.A., Herrador R., Vindigni A., Lopes M. Rad51-mediated replication fork reversal is a global response to genotoxic treatments in human cells. J. Cell Biol. 2015;208:563–579. doi: 10.1083/jcb.201406099.
    1. Mutreja K., Krietsch J., Hess J., Ursich S., Berti M., Roessler F.K., Zellweger R., Patra M., Gasser G., Lopes M. ATR-Mediated Global Fork Slowing and Reversal Assist Fork Traverse and Prevent Chromosomal Breakage at DNA Interstrand Cross-Links. Cell Rep. 2018;24:2629–2642.e5. doi: 10.1016/j.celrep.2018.08.019.
    1. Rossi S.E., Ajazi A., Carotenuto W., Foiani M., Giannattasio M. Rad53-Mediated Regulation of Rrm3 and Pif1 DNA Helicases Contributes to Prevention of Aberrant Fork Transitions under Replication Stress. Cell Rep. 2015;13:80–92. doi: 10.1016/j.celrep.2015.08.073.
    1. Couch F.B., Bansbach C.E., Driscoll R., Luzwick J.W., Glick G.G., Bétous R., Carroll C.M., Jung S.Y., Qin J., Cimprich K.A., et al. ATR phosphorylates SMARCAL1 to prevent replication fork collapse. Genes Dev. 2013;27:1610–1623. doi: 10.1101/gad.214080.113.
    1. Ragland R.L., Patel S., Rivard R.S., Smith K., Peters A.A., Bielinsky A.K., Brown E.J. RNF4 and PLK1 are required for replication fork collapse in ATR-deficient cells. Genes Dev. 2013;27:2259–2273. doi: 10.1101/gad.223180.113.
    1. Yu C., Gan H., Han J., Zhou Z.X., Jia S., Chabes A., Farrugia G., Ordog T., Zhang Z. Strand-Specific Analysis Shows Protein Binding at Replication Forks and PCNA Unloading from Lagging Strands when Forks Stall. Mol. Cell. 2014;56:551–563. doi: 10.1016/j.molcel.2014.09.017.
    1. Kubota T., Katou Y., Nakato R., Shirahige K., Donaldson A.D. Replication-Coupled PCNA Unloading by the Elg1 Complex Occurs Genome-wide and Requires Okazaki Fragment Ligation. Cell Rep. 2015;12:774–787. doi: 10.1016/j.celrep.2015.06.066.
    1. Ohashi E., Tsurimoto T. Advances in Experimental Medicine and Biology. Volume 1042. Springer; New York, NY, USA: 2017. Functions of multiple clamp and clamp-loader complexes in eukaryotic DNA replication; pp. 135–162.
    1. Errico A., Aze A., Costanzo V. Mta2 promotes Tipin-dependent maintenance of replication fork integrity. Cell Cycle. 2014;13:2120–2128. doi: 10.4161/cc.29157.
    1. Toledo L., Neelsen K.J., Lukas J. Molecular Cell Perspective Replication Catastrophe: When a Checkpoint Fails because of Exhaustion. Mol. Cell. 2017;66:735–749. doi: 10.1016/j.molcel.2017.05.001.
    1. Segurado M., Diffley J.F.X. Separate roles for the DNA damage checkpoint protein kinases in stabilizing DNA replication forks. Genes Dev. 2008;22:1816–1827. doi: 10.1101/gad.477208.
    1. Kaochar S., Shanks L., Weinert T. Checkpoint genes and Exo1 regulate nearby inverted repeat fusions that form dicentric chromosomes in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA. 2010;107:21605–21610. doi: 10.1073/pnas.1001938107.
    1. Morafraile E.C., Bugallo A., Carreira R., Fernández M., Martín-Castellanos C., Blanco M.G., Segurado M. Exo1 phosphorylation inhibits exonuclease activity and prevents fork collapse in rad53 mutants independently of the 14-3-3 proteins. Nucleic Acids Res. 2020;48:3053–3070. doi: 10.1093/nar/gkaa054.
    1. Tomimatsu N., Mukherjee B., Harris J.L., Boffo F.L., Hardebeck M.C., Potts P.R., Khanna K.K., Burma S. DNA-damage-induced degradation of EXO1 exonuclease limits DNA end resection to ensure accurate DNA repair. J. Biol. Chem. 2017;292:10779–10790. doi: 10.1074/jbc.M116.772475.
    1. Hae Y.Y., Shevchenko A., Shevchenko A., Dunphy W.G. Mcm2 is a direct substrate of ATM and ATR during DNA damage and DNA replication checkpoint responses. J. Biol. Chem. 2004;279:53353–53364. doi: 10.1074/jbc.M408026200.
    1. Shorrocks A.M.K., Jones S.E., Tsukada K., Morrow C.A., Belblidia Z., Shen J., Vendrell I., Fischer R., Kessler B.M., Blackford A.N. The Bloom syndrome complex senses RPA-coated single-stranded DNA to restart stalled replication forks. Nat. Commun. 2021;12:585. doi: 10.1038/s41467-020-20818-5.
    1. Davies S.L., North P.S., Hickson I.D. Role for BLM in replication-fork restart and suppression of origin firing after replicative stress. Nat. Struct. Mol. Biol. 2007;14:677–679. doi: 10.1038/nsmb1267.
    1. Franchitto A., Pichierri P. Bloom’s syndrome protein is required for correct relocalization of RAD50/MRE11/NBS1 complex after replication fork arrest. J. Cell Biol. 2002;157:19–30. doi: 10.1083/jcb.200110009.
    1. Godoy V.G., Jarosz D.F., Walker F.L., Simmons L.A., Walker G.C. Y-family DNA polymerases respond to DNA damage-independent inhibition of replication fork progression. EMBO J. 2006;25:868–879. doi: 10.1038/sj.emboj.7600986.
    1. Imlay J.A., Chin S.M., Linn S. Toxic DNA damage by hydrogen peroxide through the fenton reaction in vivo and in vitro. Science. 1988;240:640–642. doi: 10.1126/science.2834821.
    1. Rowe L.A., Degtyareva N., Doetsch P.W. DNA damage-induced reactive oxygen species (ROS) stress response in Saccharomyces cerevisiae. Free Radic. Biol. Med. 2008;45:1167–1177. doi: 10.1016/j.freeradbiomed.2008.07.018.
    1. Rowe L.A., Degtyareva N., Doetsch P.W. Yap1: A DNA damage responder in Saccharomyces cerevisiae. Mech. Ageing Dev. 2012;133:147–156. doi: 10.1016/j.mad.2012.03.009.
    1. Wu C.H., Jiang W., Krebs C., Stubbe J. YfaE, a ferredoxin involved in diferric-tyrosyl radical maintenance in Escherichia coli ribonucleotide reductase. Biochemistry. 2007;46:11577–11588. doi: 10.1021/bi7012454.
    1. Nakayashiki T., Mori H. Genome-Wide screening with hydroxyurea reveals a link between nonessential ribosomal proteins and reactive oxygen species production. J. Bacteriol. 2013;195:1226–1235. doi: 10.1128/JB.02145-12.
    1. Netz D.J.A., Stith C.M., Stümpfih M., Köpf G., Vogel D., Genau H.M., Stodola J.L., Lill R., Burgers P.M.J., Pierik A.J. Eukaryotic DNA polymerases require an iron-sulfur cluster for the formation of active complexes. Nat. Chem. Biol. 2011;8:125–132. doi: 10.1038/nchembio.721.
    1. Rudolf J., Makrantoni V., Ingledew W.J., Stark M.J.R., White M.F. The DNA repair helicases XPD and FancJ have essential iron-sulfur domains. Mol. Cell. 2006;26:801–808. doi: 10.1016/j.molcel.2006.07.019.
    1. Holt M.E., Salay L.E., Chazin W.J. A Polymerase With Potential: The Fe–S Cluster in Human DNA Primase. Methods Enzymol. 2017;595:361–390.
    1. 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. doi: 10.1007/s13238-015-0134-8.
    1. Kettani T., Cotton F., Gulbis B., Ferster A., Kumps A. Plasma hydroxyurea determined by gas chromatography-mass spectrometry. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2009;877:446–450. doi: 10.1016/j.jchromb.2008.12.048.
    1. Scott D.K., Neville K., Garg U. Determination of hydroxyurea in serum or plasma using gas chromatography-mass spectrometry (GC-MS) Methods Mol. Biol. 2010;603:279–287. doi: 10.1007/978-1-60761-459-3_26.
    1. Adragna N.C., Fonseca P., Lauf P.K. Hydroxyurea affects cell morphology, cation transport, and red blood cell adhesion in cultured vascular endothelial cells. Blood. 1994;83:553–560. doi: 10.1182/blood.V83.2.553.553.
    1. Cokic V.P., Smith R.D., Beleslin-Cokic B.B., Njoroge J.M., Miller J.L., Gladwin M.T., Schechter A.N. Hydroxyurea induces fetal hemoglobin by the nitric oxide–dependent activation of soluble guanylyl cyclase. J. Clin. Investig. 2003;111:231–239. doi: 10.1172/JCI200316672.
    1. Cantisani C., Kiss N., Naqeshbandi A.F., Tosti G., Tofani S., Cartoni C., Carmosino I., Cantoresi F. Nonmelanoma skin cancer associated with Hydroxyurea treatment: Overview of the literature and our own experience. Dermatol. Ther. 2019;32:e13043. doi: 10.1111/dth.13043.
    1. Contreras Castillo S., Montibus B., Rocha A., Duke W., von Meyenn F., McLornan D., Harrison C., Mullally A., Schulz R., Oakey R.J. Hydroxycarbamide effect on DNA methylation and gene expression in myeloproliferative neoplasms. Genome Res. 2021;9:gr-27006. doi: 10.1101/gr.270066.120.

Source: PubMed

3
Sottoscrivi