Next Article in Journal
Mapping Post‐Transcriptional Modifications onto Transfer Ribonucleic Acid Sequences by Liquid Chromatography Tandem Mass Spectrometry
Next Article in Special Issue
DOT1L and H3K79 Methylation in Transcription and Genomic Stability
Previous Article in Journal
m1A Post‐Transcriptional Modification in tRNAs
Previous Article in Special Issue
Emerging Roles for Ciz1 in Cell Cycle Regulation and as a Driver of Tumorigenesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Common Chemical Inductors of Replication Stress: Focus on Cell‐Based Studies

1
Institute of Molecular and Translational Medicine, Faculty of Medicine and Dentistry, Palacky University, Hnevotinska 5, Olomouc 779 00, Czech Republic
2
MRC Laboratory for Molecular Cell Biology, University College London, London WC1E 6BT, UK
*
Author to whom correspondence should be addressed.
Biomolecules 2017, 7(1), 19; https://doi.org/10.3390/biom7010019
Submission received: 25 November 2016 / Accepted: 10 February 2017 / Published: 21 February 2017
(This article belongs to the Special Issue Chromosome Maintenance)

Abstract

:
DNA replication is a highly demanding process regarding the energy and material supply and must be precisely regulated, involving multiple cellular feedbacks. The slowing down or stalling of DNA synthesis and/or replication forks is referred to as replication stress (RS). Owing to the complexity and requirements of replication, a plethora of factors may interfere and challenge the genome stability, cell survival or affect the whole organism. This review outlines chemical compounds that are known inducers of RS and commonly used in laboratory research. These compounds act on replication by direct interaction with DNA causing DNA crosslinks and bulky lesions (cisplatin), chemical interference with the metabolism of deoxyribonucleotide triphosphates (hydroxyurea), direct inhibition of the activity of replicative DNA polymerases (aphidicolin) and interference with enzymes dealing with topological DNA stress (camptothecin, etoposide). As a variety of mechanisms can induce RS, the responses of mammalian cells also vary. Here, we review the activity and mechanism of action of these compounds based on recent knowledge, accompanied by examples of induced phenotypes, cellular readouts and commonly used doses.

1. Introduction

The DNA molecule always has to keep the middle ground: it must be sufficiently rigid to maintain correct genetic information while at the same time available for ongoing processes. DNA is particularly vulnerable to insults during replication, a process where a copy of the genome is generated [1]. Replication must be tightly regulated because it is essential for genome integrity, and therefore the fate of a new cellular generation. Accurate coordination of several cellular pathways is needed to provide sufficient energy and material supply, precise timing and functional repair to overcome arising difficulties [1].
Transient slowing or disruption of replication fork (RF) progression is called replication stress (RS), which can be caused by a limitation of important factors and/or obstacles caused by intrinsic and extrinsic sources [2]. Intrinsic sources of RS involve the physiological properties of the DNA molecule, such as regions of heterochromatin structure, origin-poor regions or sites rich in some types of repetitive sequences [3,4,5]. Other intrinsic sources of RS are generated by deregulated pathways that cause over- and under-replication [6,7,8], re-replication (also known as re-duplication) [9,10], or by transcription and replication machinery collisions [9].
The most common extrinsic sources of RS are all wavelengths of ultraviolet radiation (UV) [11], ionising radiation (IR) [12] and special genotoxic chemical compounds [13] which are the main focus of this review. RS-inducing chemicals can cause a broad spectrum of DNA lesions. Alkylating agents such as methyl-methane sulfonate (MMS) [14], temozolomide and dacarbazine [15] directly modify DNA by attaching an alkyl group that presents an obstacle to RF progression. Moreover, the bifunctional alkylating compounds (e.g., mustard gas) can cause the crosslinking of guanine nucleobases [16,17] that violate the DNA structure even further [18]. Typical crosslinking agents introduce covalent bonds between nucleotides located on the same strand (intrastrand crosslinks), like cisplatin, or opposite strands (interstrand crosslink), like mitomycin C, and psoralens [18]. Crosslinks make the strands unable to uncoil and/or separate and physically block RF progression [19]. Even a small amount of unrepaired crosslinks (approx. 100–500) is reported to be lethal to a mammalian cell [20]. Furthermore, single-strand DNA breaks (SSB) and double-strand DNA breaks (DSB) represent a specific problem for ongoing replication which is well manifested by increased sensitivity of replicating cells towards radiomimetic compounds (e.g., neocarzinostatin) [21]. Other compounds do not damage the DNA structure directly but rather interfere with replication-related enzymes. Aphidicolin, an inhibitor of replicative DNA polymerases leads to uncoupling of the replicon and generation of long stretches of single-stranded DNA (ssDNA) [22]. After hydroxyurea treatment, an inhibitor of ribonucleotide reductase (RNR), the metabolism of deoxyribonucleotide triphosphates (dNTPs) is disturbed, and subsequently, the RF progression is blocked [23]. Camptothecin and etoposide, inhibitors of topoisomerase I and topoisomerase II respectively, prevent DNA unwinding and halt relaxation of torsional stress [24,25]. The most common sources of RS are illustrated in Figure 1.
Several repair pathways are essential for rapid elimination of DNA distortions and lesions introduced by the action of RS inducing compounds [26]. Removal and replacement of single base damage (e.g., oxidised and alkylated bases), is performed by base excision repair (BER) [27]. More extensive damage affecting several adjacent bases is repaired by nucleotide excision repair pathway (NER). NER is essential for repair of UV-induced damage such as cyclobutane pyrimidine dimers, or pyrimidine-pyrimidone (6-4) photoproducts and also needed for crosslinks removal caused by for example cisplatin [28]. Single-strand break repair in higher eukaryotes rely on poly(ADP-Ribose) polymerase 1 (PARP1) and X-ray repair cross complementing 1 (XRCC1) depedent recognition of the lesion, followed by end processing and ligation [29]. Double-strand breaks (DSBs) are processed by either homologous recombination (HR), or non-homologous end-joining (NHEJ). HR is active predominantly in S and G2 phases using the sister chromatid as a template for repair with high fidelity [30]. NHEJ, considered as an error prone pathway, performs DSB repair in all cell cycle stages more rapidly by direct ligation of two unprocessed (or minimally processed) DNA ends [31].
All previously described specific structures and concomitant DNA lesions can challenge the progression of RF. If the RF encounters a lesion which the replicative polymerase is unable to process as a template, it becomes stalled [32]. Stalled RFs are vulnerable structures and may undergo spontaneous collapse which leads to DSBs and genomic instability (GI) [33,34]. To avoid the harmful consequences of stalled forks, several mechanisms—DNA damage tolerance pathways (DDT)—exist to bypass the lesions and enable fork restart. One well-described process of DDT is translesion synthesis (TLS). TLS promotes “polymerase switch” from the replicative polymerase to translesion polymerases, which are able to continue replication across the lesion. TLS polymerases possess low processivity and fidelity towards the template DNA strand. Therefore TLS is often referred to as the error-prone pathway of DDT [32,34,35,36]. Among the DNA lesions which block the progression of RFs, interstrand crosslinks (ICLs) belong to the most challenging to bypass [37]. Thus, a whole group of proteins called Fanconi anaemia (FA) proteins evolved to govern the bypass and the repair of ICLs. The FA network promotes the unhooking of the ICL by specific endonucleases, bypassing the lesion by TLS polymerases or the repair by HR [5,6,7]. Patients with a defect in the FA protein family suffer from premature ageing, show increased sensitivity to DNA crosslinking agents (e.g., cisplatin, mitomycin C) and predisposition to certain types of cancers due to increased GI [38,39,40]. Although the FA pathway is involved mainly in ICL repair, it contributes more generally to initial detection of RF arrest, processing and stabilisation of the forks and regulation of TLS [41,42].
DNA damage bypass can occur in an error-free manner through the activation of the other branch of DDT, called template switching (TS). The process utilises the newly synthesised strand of the sister duplex, using it as an undamaged template. TS can be promoted either by fork regression or by strand invasion mediated by HR [34,36,43,44]. RF restart can also be achieved by firing nearby dormant replication origins or by repriming events leaving behind lesion containing single-stranded DNA (ssDNA) gaps which are subsequentially processed by DTT pathways [45,46,47,48,49,50]. Altogether, these processes ensure the rapid resumption of DNA synthesis, preventing prolonged fork stalling and the potentially deleterious effects of replication fork collapse. However, upon persisting RS, or non-functional RS response, the RF may fail to restart and collapse, most probably due to destabilised, dysfunctional or displaced components of replication machinery [1,50,51,52,53,54]. Prolonged stalled replication forks are targeted by endonucleases followed by recombination-based restart pathways [55,56].
Among the features of RS belong accumulation of long stretches of ssDNA [46,57], resulting from the uncoupled activity of DNA polymerase and progression of DNA helicase [58,59]. The persisting ssDNA is rapidly coated by replication protein A (RPA) that in turn generates the signal triggering the checkpoint response through activation of Ataxia telangiectasia Rad3-related (ATR) checkpoint kinase [60,61,62,63]. Once activated, ATR and its downstream target checkpoint kinase 1 (CHK1) help the cell to faithfully complete DNA replication upon RS [52,53,64]. In addition, ATR as the central RS response kinase contributes to the stabilisation and restart of the stalled forks even after the stress has been removed [65]. The ATR-CHK1 pathway is responsible for cell cycle inhibition, suppression of new origin firing, DNA repair and to the overall improvement of cell survival [62,66]. The role of Ataxia telangiectasia mutated (ATM), another important checkpoint kinase, upon RS conditions is not as clear and straightforward as of ATR. ATM is preferentially activated by DSBs which are generated in later stages after RS induction, mostly after the RF collapse [67,68]. There is suggested interplay between ATM and ATR during replication stress which becomes apparent under concomitant depletion of both kinases [68]. Interplay between ATM, Werner helicase (WRN) and Bloom helicase (BLM) is needed for the resolution of replication intermediates and HR repair pathway that is important for RF restart [69,70].
Chronic replication stress conditions, particularly in the absence of proper DNA repair pathway and/or non-functional checkpoint responses might result in the transfer of RS-related DNA alterations to daughter cells, inducing mutations, GI and fuelling tumourigenesis [1].
From this point of view, the RS is a strong pro-carcinogenic factor driving selective pressure for acquisition of mutations overcoming cell cycle arrest or apoptosis [71,72]. This further leads to the progression of malignant transformation and faster selection of mutations allowing development of resistance to cancer treatment [73].
However, cells typically react on the prolonged exposure to RS by triggering mechanisms leading to permanent cell cycle arrest known as cellular senescence or apoptosis [74,75] acting as a natural barrier against tumour progression [76].
Several hereditary syndromes are linked to enhanced RS and GI. The spectrum of exhibited symptoms is broad and includes premature ageing, growth retardation, neurodegeneration, immunodeficiency, cancer predisposition and others. The disorders like Seckel syndrome (deficiency in ATR kinase) [77], Ataxia telangiectasia caused (loss of ATM kinase) [78], Xeroderma pigmentosum (XP); various defects in XP protein family group) [79] are caused by aberrations in DNA damage recognition and repair enzymes [80]. Bloom and Werner syndrome (deficiency of BLM and WRN helicase, respectively) [81,82], Fanconi anaemia (FA; mutations in FA pathway proteins) [83,84], or Rothmund-Thomson syndrome (defects in RECQ like helicase 1 protein) [85,86] are related to failure of replication fork progression and restart.
In general, RS is a potent inducer of variety of hereditary and non-hereditary diseases, including the oncogenic transformation. The knowledge and understanding of the processes during RS are crucial for choosing the most efficient therapy. The in vitro-based cell studies involving models of chemical induction of RS are unique source of information about molecular interactions and undergoing mechanisms. For this review five compounds were chosen, all of them are commonly used for cell-based experiments to induce RS. Several aspects are discussed in detail: mechanism of action aimed at replication interference, proper dosing and common experimental setups. A brief overview of the medical use and important practical hints for laboratory use are also included.

2. Compounds

2.1. Cisplatin

Cisplatin (cisPt) is an inorganic platinum complex first synthesised by Italian chemist Michel Peyrone and originally known as ‘Peyrone’s chloride’ (Figure 2). The cytostatic activity of cisPt was first reported by Barnett Rosenberg and co-workers in 1965 following accidental discovery of Escherichia coli growth inhibition induced by the production of cisPt from platinum electrodes [87]. It is generally considered as a cytotoxic drug for treating cancer cells by damaging DNA and inhibiting DNA synthesis. cisPt is a neutral planar coordination complex of divalent platinum [88] with two labile chloride groups and two relatively inert amine ligands. The cis configuration is necessary for the antitumour activity [89], 3D structure of monofunctional cisPt bound to DNA structure can be found here [90].

2.1.1. Mechanism of DNA Damage Induction

The cytotoxicity of cisPt is known to be due to the formation of DNA adducts, including intrastrand (96%) and interstrand (1%) DNA crosslinks, DNA monoadduct (2%) and DNA–protein crosslinks (<1%) [91]. These structural DNA modifications block uncoiling and separation of DNA double-helix strands, events both necessary for DNA replication and transcription [92]. Inside a cell, cisPt forms an activated platinum complex, which triggers a nucleophilic substitution reaction via an attack on nucleophilic centres on purine bases of DNA, in particular, N7 positions of guanosine (65%) and adenosine residues (25%) [93]. The two reactive sites of cisPt enable the formation of the most critical crosslink between two adjacent guanines (1,2-d(GpG)), resulting in the formation of DNA intrastrand crosslinks [94]. Also, platinum can align to guanine bases on the opposite DNA strand, thus creating DNA interstrand crosslinks, present in lower percentage [95]. These cisPt crosslinks create severe local DNA lesions that are sensed by cellular proteins, inducing repair, replication bypass or triggering apoptosis [96]. Several protein families can recognise cisPt–DNA adducts, including nucleotide excision repair (NER) proteins [97], homology-directed repair proteins (HDR) [98], mismatch repair (MMR) proteins [99] and non-histone chromosomal high mobility group proteins 1 and 2 (HMG1 and HMG2) [100]. The intrastrand cisPt structural alteration stalls RNA polymerase II. It is recognised and efficiently repaired by global genome NER (GG-NER) or its transcription-coupled sub-pathway (TC-NER) [101]. The second DNA repair system predominantly involved in coping with cisPt–DNA adducts is error-free HDR, which removes DNA DSBs remaining after cisPt adduct removal [98]. In contrast to the previously mentioned repair pathways that increase cell viability, MMR proteins have been shown to be essential for cisPt-mediated cytotoxicity [99]. cisPt is reported to enhance interactions between MMR proteins MLH1/PMS2 (MutL homolog 1/PMS1 homolog 2, MMR component) and p73, triggering apoptosis [102]. Therefore, mutations in MMR genes are known to be associated with cisPt resistance [103]. HMG1 and HMG2 recognise intrastrand DNA adducts between adjacent guanines, affecting cell cycle events and subsequently inducing apoptosis [100].
In addition to the previously mentioned repair proteins, specialised translesion DNA polymerase eta (η) can be loaded onto sites of cisPt–DNA adducts promoting TLS repair pathway [104]. cisPt also induces dose-dependent reactive oxygen species (ROS), which are responsible for the severe side effects of platinum-based therapy, including nephrotoxicity and hepatotoxicity [105]. When overwhelming the reduction capacity of the cell, cisPt-induced ROS might lead to lipid peroxidation, oxidative DNA damage, altered signal transduction pathway and calcium homoeostasis failure [105]. Extensive unrepaired cisPt-induced DNA damage can proceed to apoptotic cell death mediated by various signal transduction pathways, including calcium signalling [106], death receptor signalling [107] and activation of mitochondrial pathways [108]. At least two main pathways have been proposed to mediate cisPt-induced apoptosis in vitro. One involves the critical tumour suppressor protein p53 directly binding to cisPt-modified DNA [109] and promoting apoptosis via several mechanisms. p53 binds and counteracts the anti-apoptotic B-cell lymphoma-extra large (Bcl-xL) [110], contributes to inactivation of nutrient sensor AMP-kinase (AMPK) [111], activates caspase-6 and -7 [112] and the pro-apoptotic Bcl-2 family member PUMAα in renal tubular cells [113]. However, the role of p53 in response to cisPt seems to be controversial, as it has been described to contribute to cisPt cytotoxicity [114] and also to be involved in cisPt resistance in different cancer models [115]. The other cisPt-induced apoptotic pathway is mediated via a pro-apoptotic member of the p53 family, p73. cisPt has been shown to induce p73 in several cancer cell lines [116], which cooperates with the MMR system and c-Abl tyrosine kinase, known to be involved in DNA damage-induced apoptosis [117]. In response to cisPt, c-Abl phosphorylates p73, making it stable [118], and increases its pro-apoptotic function by binding transcription coactivator p300, which triggers transcription of pro-apoptotic genes [119]. Moreover, p73 forms a complex with c-Jun N-terminal kinase/stress-activated protein kinase (JNK), leading to cisPt-induced apoptosis [120]. Intrinsic signaling pathways involved in cisPt driven apoptosis include Akt [121], protein kinase C [122,123], and mitogen activated protein kinases—MAPK (e.g., extracellular signal-regulated kinases; ERK) [124,125,126], JNK [127,128,129] and p38 [130].

2.1.2. Other Effects

Besides DNA, the primary target of cisPt in cells, there is some evidence for the involvement of non-DNA targets in cisPt cytotoxicity [131]. cisPt interacts with phospholipids and phosphatidylserine in membranes [132], disrupts the cytoskeleton and alters the polymerization of actin, probably due to conformational changes resulting from the formation of Pt–S bonds [133]. MicroRNAs (miR), which play a role in posttranscriptional gene silencing, have been shown to be involved in the modulation of cisPt resistance-related pathways in different cancer models. miR-378 was shown to reverse resistance to cisPt in lung adenocarcinoma cells [134], whereas miR-27a was shown to be upregulated in a multidrug resistant ovarian cancer cell line, contributing to cisPt resistance [135]. miR-21 increases the cisPt sensitivity of osteosarcoma-derived cells [136]. For references to particular studies using cisPt, refer to Table 1.

2.1.3. Solubility

cisPt (molecular weight (MW) 300.05 g/mol) is water soluble at 2530 mg/L (at 25 °C), saline solution with a high chloride concentration (approx. 154 mmol/L) is recommended. In the absence of chloride, the cisPt chloride leaving group becomes aquated, replacing the chloride ligand with water and generating a mixture of species with increased reactivity and altered cytotoxicity [150,151]. Commonly used solutions for laboratory use are aqueous-based solutions in 0.9% NaCl (0.5 mg/mL), pH 3.5–5. Dissolved cisPt may degrade over a short time, the storage of aliquots is not recommended. However, the stability at −20 °C in the dark is reported to be 14 days. Solutions (in 2 mM phosphate buffered saline buffer with chloride concentration 140 mmol/L) stored at 4 °C should be stable for 7–14 days [152]. Undiluted cisPt is stable in the dark at 2–8 °C for several months [121,153]. Dimethyl sulfoxide (DMSO) can also be used for cisPt dilution, however it is not recommended. The nucleophilic sulphur can displace cisPt ligands, affecting the stability and reducing cisPt cytotoxicity [154]. DMSO introduced in combination studies with cisPt does not affect its activity [152].

2.1.4. Medical Use

Following the start of clinical trials in 1971, cisPt, marketed as Platinol (Bristol-Myers Squibb, New York, USA), was approved for use in ovarian and testicular cancer by the Food and Drug Administration (FDA) in 1979 [155]. cisPt is considered one of the most commonly used chemotherapy drugs for treating a wide range of malignancies, including head and neck, bladder, oesophagal, gastric and small cell lung cancer [156,157]. Moreover, cisPt has been shown to treat Hodgkin’s [158] and non-Hodgkin’s lymphomas [159], neuroblastoma [160], sarcomas [161], multiple myelomas [162], melanoma [163], and mesothelioma [164]. cisPt can reach concentrations of up to 10 μg/mL in human plasma [165]. cisPt is administrated either as a single agent or, in the main cases, in combination with other cytostatics (e.g., bleomycin, vinblastine, cyclophosphamide) or radiotherapy for the treatment of a variety of tumours, e.g., cervical carcinoma [153]. The most important reported side effect is nephrotoxicity, due to preferential accumulation and persistence of cisPt in the kidney [166], later ototoxicity and bone marrow depression. Pharmacokinetic and pharmacodynamic studies have shown that a maximal steady state cisPt plasma concentration of between 1.5 and 2 μg/mL has the most effective chemotherapeutical effect with minimal adverse nephrotoxicity [167]. Many cancers initially responding to cisPt treatment could become later resistant. Mechanisms involved in the development of cisPt resistance include changes in cellular uptake, drug efflux, drug inactivation by increased levels of cellular thiols, processing of cisPt-induced damage by increased NER and decreased MMR activity and inhibition of apoptosis [99,168]. To boost platinum drug cytotoxicity, overcome its resistance and achieve a synergistic effect, new platinum-based drugs, as well as their combinatorial therapy with other antineoplastic agents were developed for cancer treatment [169]. Besides of cisPt derivatives as carboplatin and oxaliplatin, are currently being used in the clinical practice, while nedaplatin, lobaplatin and nedaplatin acquired limited approval in clinical use [170,171]. Recent discoveries described the combination of cisPt with PARP inhibitor olaparib targeting DNA repair to acts synergistically in several non-small cell lung carcinoma cell lines [172]. This combinatorial therapy can be promising especially in patients with advanced breast and ovarian cancer-bearing BRCA1/2 (breast cancer 1/2) mutations [173].

2.1.5. Summary

cisPt is used in vitro in concentration range approx. 0.5–300 μM. The levels in human plasma can reach up to 10 μg/mL (33 μM) which should be beared in mind when interpreting in vitro data. Continuous treatment, or longer incubation time, or high cisPt concentration of 20 mg/mL lead to complete inhibition of DNA synthesis [174]. The concentration range of 15–30 μM results in a block of DNA replication and transcription and triggers DNA damage response (DDR) signalization through ATM-CHK2, ATR-CHK1 DDR pathways resulting in p53-p21 driven cell cycle arrest or p53-mediated cell apoptosis [141,142,143,144]. However, in some cell lines also the synthesis of anti-apoptotic protein Bcl-2 was reported [143]. cisPt is in the majority of cell lines induces apoptosis above the concentration of approx. 2 μM [139,141,142,146]. cisPt block DNA replication [139,140,146] and inhibits RNA synthesis [140,175,176] and also influences the mitochondrial DNA synthesis and metabolism [147]. As a commonly used drug in clinics, many in vitro experiments have been conducted to address problems arising during treatment. Especially, the study of mechanisms underlying drug resistance [177], causes of toxic side effects [178], enhancement of synergistic effects [179] and ways how to improve drug delivery systems [180]. cisPt massively triggers the TLS repair pathways; defective FA proteins sensitise the cells towards this compound [181], defective MMR proteins establish cisPt resistance [103,182].

2.2. Aphidicolin

Aphidicolin (APH) is a tetracyclin diterpenoid antibiotic isolated from Nigrospora sphaerica (Figure 3) which interferes with DNA replication by inhibiting DNA polymerases α, ε and δ [183]. Specifically, only cells in S phase are affected, whereas cells in other phases of the cell cycle are left to continue until the G1/S checkpoint, where they accumulate [184].

2.2.1. Mechanism of DNA Damage Induction

APH binds to the active site of DNA polymerase α and rotates the template guanine, selectively blocking deoxycytidine triphosphate (dCTP) incorporation [185]. DNA polymerase α interacts with APH by its C18-binding OH group, APH forms a transient complex with polymerase and DNA [183]. The effect of APH on cell cultures is reversible if the cells are treated for no longer than 2 generations [186]. The exonuclease activity of APH-responding polymerases is only mildly affected, even at concentrations completely blocking the polymerase activity [183]. However, in the cell nucleus, the exonuclease activity is usually not retained because ternary complex APH–polymerase–DNA is formed and blocks the enzyme [183]; 3D structure of the complex can be found here [187].
Mechanistically, APH compromises the function of DNA polymerase, while helicase proceeds regularly (so called uncoupled/disconnected replicon), which leads to the generation of long stretches of single-stranded DNA [188]. The disconnected replicon is vulnerable structure prone for breakage preferentially at the so-called common fragile sites (CFSs) (also referred to as CFS expression) [189]. CFSs are specific genomic loci conserved in mammals generally prone to instability upon RS [190]. CFS expression is also common in precancerous and cancerous lesions [76]. Moreover, a causative role of CFS’s in cancer development has been suggested [191]. APH reproducibly causes damage at the same sites, and thus low doses of APH are used to define APH-inducible CFSs, of which there are over 80 described in the human genome [22,192]. Other CFS inducers (hydroxyurea, camptothecin, hypoxia and folate deficiency) are not so specific, nor efficient as APH [193,194]. Importantly, APH efficiently induces CFS expression only when the rate of polymerase is slowed down but not completely blocked. The optimum concentration range usually spans 0.1–1 μM [195] (and refer to Table 2). Apart from disconnected replicon, there might be other explanations for the extraordinary potency of APH to induce CFS-associated genomic instability. First, APH has been shown to increase the number of R-loops within certain CFSs, thus inducing replication/transcription collisions [196]. However, the mechanistic relationship between APH and increased R-loop formation is not clear. Second, re-licensing of replication origins is typical feature of oncogenic genetic backgrounds which are very prone to CFS expression. In such situations the CFS expression is explained as a result of DNA re-replication and subsequent collision of re-replicating forks within CFSs [10,197]. This phenomenon was studied in detail in yeasts at replication slow zones (analogs to CFSs in mammals) [198]. It is not clear whether the same re-licensing process is induced also by APH, however re-duplication would explain the reported APH-induced amplifications [191,199].
Prolonged treatment with low doses of APH induces cellular senescence response [74]. Interestingly, the most efficient doses were found to span the same range as doses used for CFS expression, which implies that CFSs might play a causative role in this process. Moreover, oncogene-induced senescence also displays increased CFSs-associated instability [10,197]. These phenotypical similarities between oncogenic stress and low doses of APH make this drug a good candidate for studying cellular processes in early stages of malignant transformation.

2.2.2. Other Effects

APH is a very specific DNA polymerase inhibitor, APH does not interact with mitochondrial DNA polymerases [186] nor proteins [200], DNA, RNA, metabolic intermediates, nor nucleic acid precursor synthesis [184,200,201]. Contradictory results have been obtained regarding the effect of APH on DNA repair synthesis (DRS). According to a radiography method, APH does not influence DRS [200], although when DRS was induced by tumor necrosis factor (TNF) or UV irradiation, APH was observed to inhibit the process [202,203]. For references to particular studies using APH, refer to Table 2.

2.2.3. Solubility

APH (MW 338.48 g/mol) is soluble in DMSO (up to 10 mg/mL), ethanol (up to 1 mg/mL) and methanol (freely), not soluble in water. The stability of the powder is 3 years at 2–8 °C, ethanol solution for a week at 2–8 °C, DMSO solution for 6 weeks at −20 °C [218].

2.2.4. Medical Use

APH has limited use in clinical practice owing to its low solubility. Only APH-glycinate has so far been tested in clinical trial phase I. However, fast clearance from human plasma (no drug observed after 6–8 h of APH administration) and no anti-tumour activity was observed. Its use as a single agent or even in combination with other cytostatics is no longer being considered [219]. APH is metabolised by cytochrome P-450 dependent degradation [220]. APH and its derivatives are considered as potential therapeutics for parasitic diseases, e.g., Chagas disease [221].

2.2.5. Summary

APH is used for in vitro studies in concentration range approx. 0.01 μM to 0.2 mM. APH is mainly used for cell-based experiments involving CFS expression [222], cell cycle synchronization [223], replication fork stability and restart studies [224] and for cellular senescence induction [74]. The threshold between replication fork stalling and slowing down is around 1 μM. Upon higher concentrations (5 μM–0.2 mM) APH was reported to stall the DNA polymerase, leading to S phase arrest. Upon lower concentrations, when the DNA polymerases are just slowed down, CFS expression can be observed. Usually, longer incubation times (approx. one population doubling) are used, so more cells within the population are affected. APH treatment causes a significant amount of DNA damage, leading to rapid ATR kinase activation. In the case of longer APH treatment also ATM is activated probably as a consequence of DSB formed within the stalled replication forks [207]. Prolonged APH incubation in the range of days up to weeks at low concentrations (0.2–1 μM) induces cellular senescence [74].

2.3. Hydroxyurea

Hydroxyurea (HU) was first synthesised in the 19th century (Figure 4) and inhibits the incorporation of nucleotides by interfering with the enzyme ribonucleotide reductase (RNR) [225]. RNR converts nucleotide di- and tri-phosphates to deoxynucleotide di- and tri-phosphates, which is the rate-limiting step in nucleotide synthesis [226]. Without proper levels of dNTPs, DNA cannot be correctly replicated nor repaired [227].

2.3.1. Mechanism of DNA Damage Induction

RNR is a large tetrameric enzyme comprising two R1 subunits and two small regulatory subunits R2 [228]. HU scavenges the tyrosyl radical of the R2 subunit which inactivates the RNR enzymatic activity [226]. Complete inhibition of RNR has been observed within 10 min after treatment with 0.1 mM HU and within 5 min after 3 mM of HU in murine 3T6 cells [229]. Consequently DNA synthesis is inhibited, selectively stopping the cells in S phase [230]. The inhibition is caused alterations in the dNTP pools. Each type of dNTP is affected in a different way. For example, after 280–560 μM HU treatment for 60 min, the dTTP pool was found to increase by 50%, whereas the dCTP pool is decreased by 50% [231]. HU slows down the initiation of replication and also the progression of replication forks. Moreover, after stopping the production of dNTPs, DNA repair and mitochondrial DNA synthesis are affected in all cells, regardless of the cell cycle stage [227]. HU treatment greatly affects the choice of replication origins and origin spacing in mammalian cells [232]. Although the mechanism of DNA damage induction may look similar to that for APH, HU induces a different spectrum of fragile sites, called early replicating fragile sites (ERFs) [233]. ERFs are also induced by c-Myc expression [11,12]. It was also reported that 10 μg/mL APH [234] (concentration that stalls the replication fork progression) leads to the generation of several kilobases long unwound DNA; however, HU treatment can generate only up to 100–200 nt long ssDNA [235].

2.3.2. Other Effects

HU induces copy number variants (CNVs) with similar frequency and size distribution as APH [236]. It was reported for yeast cells, that HU alters Fe–S centres, enzyme cofactors catalysing oxidation-reduction reactions, which interferes with various metabolic enzymes and affects the redox balance of cells. Similar mechanism is proposed also for mammalian cells [237].
HU has been negatively tested for mutagenicity, measured by single nucleotide variation (SNV) and insertion/deletion frequency [238]. On the other hand, low doses of HU have been reported to induce DNA damage [239]. Therefore, it is possible that the compound possesses some pro-mutagenic potential (see also below). For references to particular studies using HU, refer to Table 3.

2.3.3. Solubility

HU (MW 76.05 g/mol) is freely soluble in water at 100 mg/mL, soluble also in DMSO. The powder is stable at 4 °C for 12 months. Solutions are stable for 1 month at −20 °C (after defrosting, equilibration is recommended for 1 h at room temperature. It is recommended to prepare fresh solutions before use. HU decomposes in the presence of moisture; therefore, it is recommended that it is stored in air-tight containers in a dry atmosphere [259].

2.3.4. Medical Use

HU is a commonly used medicine first approved by the FDA for the treatment of neoplastic disorders in the 1960s [260]. Common plasma levels of HU range 100–200 μM [261]. It is used for the treatment of sickle cell disease, essential thrombocytosis [262], myeloproliferative disorders and psoriasis [260] and is commonly indicated as a cytoreductive treatment in polycythemia vera [263] and others. Synergistic effects have been reported when it is used in combination with antiretroviral pills [264] and also in indicated cases with radiotherapy [265]. HU may be used as an anti-retroviral agent, especially in HIV (human immunodeficiency virus) patients. HU may cause myelofibrosis development with increased time of use and AML/MDS syndrome (acute myeloid leukaemia/myelodysplastic syndrome) [266]. Adverse side-effects have been observed, mainly myelosuppression [267]. A 17-year follow-up study of 299 patients treated with HU as a long-term therapy showed no difference in the incidence of complications such as stroke, renal disease, hepatic disease, malignancy or sepsis [268], suggesting that HU is well-tolerated. However, CNVs are generated at therapeutic doses of HU, and data from reproductive studies and studies on subsequent generations have so far been rather limited [236,268].

2.3.5. Summary

HU is used in vitro approx. in the range 2 mΜ–5 mM. The most commonly used concentrations are around 2 mM. HU is used for cell cycle synchronization [269], replication fork stability studies [249,252], studies of recovery mechanisms after the release of RS [242] and checkpoint responses [241]. Lower concentrations are used for RS induction [254], induction of senescence [74], apoptosis [257], and repair pathways induction [217]. HU reaches plasma concentrations around 0.1 mM; this should be bear in mind when interpreting the data for clinical relevance [261]. The MRN (Mre11-Rad50-Nbs1) complex members Mre11 (Meiotic recombination 11) and Nbs1 (Nijmegean breakage syndrome 1) are required for efficient recovery of replication after treatment with replication stalling agents such as hydroxyurea [12]. HU causes rapid generation of ssDNA as indicated by RPA loading 40 min after treatment [270]. Subsequently, ATR-CHK1 signalling is activated, and HR repair pathway is induced.
Cells deficient in XRCC2 or other homologous recombination components exhibit hypersensitivity to HU [271]. It was reported that for hamster V79 cells, low concentrations of HU (5–10 μM) mimics the replication dynamics of untreated HR deficient cells [217]. Cellular senescence after long term replication stress caused by HU is dependent on p53-p21 signalling pathway and independent of p16 [74]. HU influences mutiple cellular pathways, e.g., JNK pathway, mitochondrial and peroxisome biogenesis, expression of several heat shock response proteins, autophagy pathways stimulation (beclin-1 expression), hemoglobin type F induction (in sickle cell disease, β-thalasemmia patients), etc. [272]. There are several cell lines that response to HU treatment in a specific manner, e.g., K562 cell line undergoes differentiation [253], T-cells activation is decreased [264], the morphology of vascular endothelial cells is affected [273].

2.4. Camptothecin

Camptothecin (CPT) is a pentacyclic quinoline alkaloid first isolated from the Chinese tree Camptotheca acuminata (Nyssaceae) by Wall et al. [274] (Figure 5). CPT has a unique intracellular target, topoisomerase I (TopoI), a nuclear enzyme that reduces the torsional stress of supercoiled DNA [24]. This activity enables specific regions of DNA to become sufficiently exposed and relaxed to facilitate essential cellular processes, such as DNA replication, recombination and transcription [275].

2.4.1. Mechanism of DNA damage induction

TopoI binds covalently to double-stranded DNA through a reversible transesterification reaction, generating a SSB [276], 3D structure can be found here [277]. This so-called TopoI–DNA cleavage complex (Top1cc) facilitates the relaxation of torsional strain in supercoiled DNA, either by allowing passage of the intact single strand through the nick or by free rotation of the DNA around the uncleaved strand [278]. CPT covalently and reversibly stabilises the normally transient DNA Top1cc by inhibiting religation of the scissile strand, thereby prolonging the half-life of Top1cc and increasing the number of DNA SSBs [279,280]. Moreover, trapping of the enzyme on the DNA leads to rapid depletion of the TopoI pool [281]. The effect of CPT is readily reversible after removal of the drug. However, prolonged stabilisation of Top1cc can cause multiple problems. Firstly, failure to relieve supercoiling generated by such processes as transcription and replication can lead to RS by creating torsional strain within the DNA [279,281,282]. Furthermore, the collision of the RF with the ternary drug-enzyme-DNA complex generates DSBs with serious cellular consequences, including cell death [283,284].
Because ongoing DNA synthesis is important for CPT-induced cytotoxicity, CPT is considered an S phase-specific drug. The repair of CPT-induced DSBs involves multiple DNA damage repair proteins. Recent studies have highlighted that functional cooperation between BRCA2, FANCD2, RAD18 and RAD51 proteins are essential for repair of replication-associated DSBs through HR. Loss of any of these proteins causes disruption of HR repair, chromosomal aberrations and sensitization of cells to CPT [285]. A close link between CPT and HR has also been demonstrated in experiments measuring sister chromatid exchange events (SCEs), which are common consequence of elevated HR repair process and found to be induced by low doses of CPT [270]. CPT is applied in early S phase cells for triggering G2 arrest accompanied by blockage of the p34cdc2/cyclin B complex, a consequence of either DNA breakage, the arrest of the replication fork or both [286]. In addition, CPT driven TopoI–DNA cleavable complex and associated strand breaks were shown to increase transcription of the c-Jun early response gene, which occurs in association with internucleosomal DNA fragmentation, a characteristic mark of apoptosis [287]. Noncytotoxic concentrations of CPT can induce the differentiation of human leukaemia cells [288], and an antiangiogenic effect is suggested [289,290]. Interestingly, when used in combined treatment with APH, CPT reduces the APH-induced RPA (an indicator of ssDNA) signal and has a rescuing effect on CFS expression [291]. For references to particular studies using CPT, refer to Table 4.

2.4.2. Solubility

CPT (MW 348.35 g/mol) is soluble in DMSO (up to 10 mg/mL), methanol (40 mg/mL), 0.1 N sodium hydroxide (50 mg/mL) or acetic acid, insoluble in water. At higher concentrations, heating is required to dissolve the product completely (approx. 10 min at 95 °C), but some precipitation occurs upon cooling to room temperature [301].

2.4.3. Medical Use

CPT cannot be used in clinical practice because of its poor solubility in aqueous solutions, instability and toxicity, but modifications at selected sites have improved the pharmacologic and activity profile [283]. Currently, three water-soluble CPT-derivates, i.e., irinotecan (CPT-11), topotecan (TPT) and belotecan (CKD-602), are available for cancer therapy. However, despite their selectivity for TopoI and unique mechanism of action, they all have critical limitations. In particular, they become inactivated against TopoI within minutes at physiological pH due to spontaneous lactone E-ring opening [302] and diffuse rapidly from the TopoI–DNA cleavage complex due to their noncovalent binding. To overcome these problems, five-membered E-ring CPT-keto non-lactone analogues S38809 and S39625 have been synthesised and selected for advanced preclinical development based on their promising activity in tumour models. Their chemical stability and ability to produce high levels of persistent Top1cc makes them useful candidates for future treatment [303].

2.4.4. Summary

Camptothecin is used in concentration range 2.5 nM up to 20 μM. CPT is a potent DSBs inducer in a wide concentration range, approx. 10 nM–10 μM. Upon higher concentration (20 μM–10 μM), CPT was reported to be cytotoxic, increasing cell apoptosis via DNA fragmentation predominantly in S phase cells with ongoing DNA synthesis [292,293]. The most frequently used concentration of 1 μM CPT was shown to block DNA synthesis and induce DSBs resulting from the collision of RF due to prolonged stabilisation of TopoI DNA cleavage complex. The main implication of lower CPT concentrations is the induction of replication fork slowing and reversal, as a rapid response to TopoI inhibition is the increase in topological stress of DNA locally [300]. CPT activates predominantly ATR-CHK1 and ATM-CHK2 signalling, and leading to G2 checkpoint arrest [300]. Even at low doses of CPT HR repair pathway is triggered.

2.5. Etoposide

Etoposide (ETP) is a derivative of podophyllotoxin first synthetised in 1966 and approved for treatment as an antineoplastic agent in 1983 [304]. ETP structure comprises of polycyclic A–D rings, an E-ring and aglycone core (Figure 6). ETP compromises the proper function of the enzyme topoisomerase II (TopoII), 3D structure can be found here [305]. TopoII performs cleavage of both strands of a DNA duplex and enables passage of a second intact duplex through the transient break, ATP is used to power the strand transition [306]. As a result, relaxation, unknotting and decatenation of DNA are achieved enabling processes like replication and transcription [25].

2.5.1. Mechanism of DNA Damage Induction

Two modes of action were suggested for ETP to interfere with TopoII [25]. As a poison, it stabilises TopoII:DNA complexes, whereas as an inhibitor ETP interacts with the catalytic site of TopoII, decreasing the number of active cleavage complexes [307]. ETP acts as a poison by stabilizing the cleavage complex of TopoII via decoupling the key catalytic residues, thus preventing the religation of cleaved DNA ends [308]. As a result, the number of TopoII-associated DNA breaks are increased [309]. ETP’s A, B and D-rings mediate the drug-enzyme interaction, whereas the aglycon core binds to DNA [262,308]. E–ring substituents are important for ETP activity but do not contribute to ETP-enzyme binding [310]. ETP is metabolised by cytochrome P3A4 (CYP3A4) to two metabolites, ETP-quinone and ETP-catechol. Both active against the TopoII enzyme. ETP-quinone is approx. 100× more efficient at inhibiting TopoII than ETP. ETP-quinone can block binding of the enzyme to DNA by stabilisation of the N-terminal clamp [307]. In cases where the enzyme still binds to DNA, the metabolite can stabilise the enzyme:DNA complex by inhibiting the religation step thus leading to higher levels of DSBs [307]. The ETP-catechol metabolite works similarly to the parent compound but can also be oxidised to the quinone [311]. ETP induces DSBs directly in all phases of the cell cycle, as observed by γH2AX foci formation (a marker of DSBs) [312,313]. ETP does not require S-phase to induce damage, but ongoing replication enhances its cytotoxic effect [314]. ETP causes disassembly of replication factories (sites of ongoing replication), as measured by the distribution of proliferating cell nucelar antigen protein (PCNA) [315]. Moreover, the cytotoxic effect of ETP is partially reduced by inhibitors of DNA synthesis, such as APH and HU [316]. There are two isoforms of the TopoII enzyme in mammals, called TopoIIα and TopoIIβ, sharing 68% homology [317]. TopoIIα activity is upregulated during cell cycle progression, peaks in mitosis and is essential for proliferating cells [318]. TopoIIβ is needed during transcription and DNA repair, and its levels are more stable during the cell cycle [319]. ETP is not selective between these two paralogs, and the inhibition of TopoIIβ is believed to be the reason for ETP therapy-related secondary malignancies [320]. TopoIIα seems to be a better target for therapy. Therefore, new compounds and analogues of ETP have been synthesised to be selective only for TopoIIα [321].

2.5.2. Other Effects

A strong mutagenic effect has been measured for ETP in mammalian cells by several assays, e.g., HPRT assay (hypoxanthine phosphoribosyl transferase), SCE and detection of mutations at the locus of the adenosine kinase gene [322]. In prokaryotic organisms (E. coli, Salmonella typhimurium), no significant genotoxic effect was observed [322]. For references to particular studies using ETP, refer to Table 5.

2.5.3. Solubility

ETP (MW 588.56 g/mol) is soluble in organic solvents (ethanol, methanol, DMSO), poorly soluble in water. It is recommended that stock solutions in organic solvents be diluted so 0.1% organic solvent is present in the final solution. The stability in aqueous solution is best at pH 4–5, but it can be improved by adding polysorbate 80 (Tween80), polyethylene glycol 300, citric acid and alcohol. ETP is unstable under oxidative conditions [338]. Under acidic conditions (pH < 4), the glycosidic linkage and lactone ring are hydrolysed, whereas, under basic conditions (pH > 6), cis-lactone epimers are formed [304]. Aqueous solutions are stable for several hours, depending on the concentration of the solution but irrespective of the temperature. ETP is sensitive to UV irradiation, both in solution and as a powder [338].

2.5.4. Medical Use

According to pharmacokinetic studies, plasma levels of ETP peak at concentrations of 20–70 μM [339]. ETP is approved for the treatment of refractory testicular tumors and small cell lung cancer. Various chemical modifications with potential higher efficacy have also been tested for clinical use, e.g., 4′-phosphorylation or 4′-propyl carboxy derivatives [340]. In the field of so-called personalised medicine, combined subsequent treatment of ETP and cisPt has been shown to be beneficial for patients suffering from ERCC1-incompetent lung adenocarcinoma [341]. ETP is reported to cause therapy-related leukaemias [320] and specific chromosomal translocations. Chromosomal rearrangements at the 11q23 chromosome band were found in patients and seemed to be related to the CYP3A4 metabolic conversion of ETP [342]. In mouse embryonic stem cells, an increase in fusion chimeric products was observed at a 1.5 kb “hot spot” between exons 9 and 11 (analogous region to MLL (mixed lineage leukaemia) breakpoint cluster in human leukaemia) [343]. MLL gene encodes lysine (K)-specific histone methyltransferase 2A therefore influencing histone methylation and gene expression [344]. Leukaemogenic MLL translocations lead to expression of MLL fusion proteins. Patients with such translocations exhibit poor prognosis [345]. MLL fusion proteins are efficient in transforming the hematopoetic cells into leukaemia stem cells [346]. Many studies have attempted to solve the adverse side effects of ETP treatment and understand the underlying molecular mechanisms, e.g., multi-drug resistance [347], or unwanted toxicity [348]. The search for compounds that may improve ETP treatment usually starts with cell-based experiments, e.g., protective compounds shielding healthy cells [349], compounds selectively enhancing ETP toxicity [350] or targeted delivery [351].

2.5.5. Summary

ETP is commonly used for the induction of apoptosis [352]. Indeed, several studies reported that higher doses of the compound (25–100 µM) activate apoptosis, mostly in a manner dependent on p53 [325,326,327,329]. Prolonged treatment at lower concentrations of ETP can also lead to induction of the p53 pathway, cell cycle arrest, senescence and apoptosis [145,325,330,335,337]. ETP induces the formation of irreversible DNA–TopoII cleavage complexes (TopoIIcc) and DNA damage regardless of concentration or incubation time [323,324,329,330,331,332,334,353]. The initial displacement of TopoIIcc requires the coordinated action of several processes, such as cleavage by the 5’-tyrosyl DNA phosphodiesterase (TTRAP) and proteasome-dependent degradation of TopoII [354,355]. Furthermore, the MRN complex, CtIP (RBBP8 protein) and BRCA1 play a critical role in the removal of such DNA-protein adducts [356]. The remaining DNA lesions are often referred as DSBs, which are accompanied by the activation of ATM-mediated signalling or repair pathways, usually quantified by the formation of γH2AX [323,324,329,330,331,332]. However, several studies argue against the ability of ETP to primarily induce DSBs, showing that majority of the DNA lesions formed upon ETP treatment are SSBs [323,329]. Despite the discrepancy, pathways engaged in DSB repair are activated after the exposure to the drug, and among them, NHEJ is seemingly predominant [329,356,357,358]. ETP used in relatively high concentration (20–25 µM) might lead to persistent or irreparable DSB formation [329,331,332].

3. Conclusions

Replication stress is a significant contributor to genomic instability, a major factor for the conservation of mutations [1], relevant promoter of tumourigenesis [8] and also one of the main features of cancer cells [76]. Owing to its complexity, replication can be disturbed by multiple mechanisms. In this review, we focused on several compounds known to be RS inducers and often used in cell-based assays. Some of the compounds have been shown to be effective in cancer treatment. Importantly, the chemicals have been primarily chosen to cover various mechanisms of action, resulting in different treatment-induced phenotypes resembling those of RS in carcinogenesis. Induction of RS in vitro, e.g., by chemicals inducing DNA damage, is a crucial research tool. Precise knowledge about the mechanism of DNA damage induction and cellular pathways involved in the RS response is particularly important for the development of appropriate cellular assays for investigating carcinogenesis and cancer treatment. The above-mentioned publications in separate compound-related tables were chosen to help with the practical aspects of such assay design. Dose and time-dependent effects related to the genetic backgrounds (i.e., dependent on the cell line used) and proper readout are important issues for experiment design. Moreover, other practical information has been included so that readers can use this review as a brief guide for choosing an appropriate model and dose scheme for cell-based studies.

Acknowledgments

The work was supported by Grant Agency of the Czech Republic 13-17555S, Internal Grant of Palacky University IGA-LF-2016-030, the Norwegian Financial Mechanism CZ09 (Project PHOSCAN 7F14061).

Author Contributions

E.V. wrote the chapter 2.3, contributed to chapters 2.2 and 2.5, prepared the figures. K.C. wrote the chapters 2.1 and 2.4, Z.T. contributed to chapter 2.5, M.M. contributed to chapter 2.2. All authors participated on the introduction part and revised the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zeman, M.K.; Cimprich, K.A. Causes and consequences of replication stress. Nat. Cell Biol. 2014, 16, 2–9. [Google Scholar] [CrossRef] [PubMed]
  2. Burhans, W.C.; Weinberger, M. DNA replication stress, genome instability and aging. Nucleic Acids Res. 2007, 35, 7545–7556. [Google Scholar] [CrossRef] [PubMed]
  3. 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. [Google Scholar] [CrossRef] [PubMed]
  4. Gelot, C.; Magdalou, I.; Lopez, B.S. Replication stress in Mammalian cells and its consequences for mitosis. Genes 2015, 6, 267–298. [Google Scholar] [CrossRef] [PubMed]
  5. Krasilnikova, M.M.; Mirkin, S.M. Replication stalling at Friedreich’s ataxia (GAA)n repeats in vivo. Mol. Cell. Biol. 2004, 24, 2286–2295. [Google Scholar] [CrossRef] [PubMed]
  6. Neelsen, K.J.; Zanini, I.M.Y.; Mijic, S.; Herrador, R.; Zellweger, R.; Ray Chaudhuri, A.; Creavin, K.D.; Blow, J.J.; Lopes, M. Deregulated origin licensing leads to chromosomal breaks by rereplication of a gapped DNA template. Genes Dev. 2013, 27, 2537–2542. [Google Scholar] [CrossRef] [PubMed]
  7. Porter, A.C. Preventing DNA over-replication: A Cdk perspective. Cell Div. 2008, 3, 3. [Google Scholar] [CrossRef] [PubMed]
  8. Burrell, R.A.; McClelland, S.E.; Endesfelder, D.; Groth, P.; Weller, M.-C.; Shaikh, N.; Domingo, E.; Kanu, N.; Dewhurst, S.M.; Gronroos, E.; et al. Replication stress links structural and numerical cancer chromosomal instability. Nature 2013, 494, 492–496. [Google Scholar] [CrossRef] [PubMed]
  9. Liu, B.; Alberts, B.M. Head-on collision between a DNA replication apparatus and RNA polymerase transcription complex. Science 1995, 267, 1131–1137. [Google Scholar] [CrossRef] [PubMed]
  10. Bartkova, J.; Rezaei, N.; Liontos, M.; Karakaidos, P.; Kletsas, D.; Issaeva, N.; Vassiliou, L.-V.F.; Kolettas, E.; Niforou, K.; Zoumpourlis, V.C.; et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 2006, 444, 633–637. [Google Scholar] [CrossRef] [PubMed]
  11. Vallerga, M.B.; Mansilla, S.F.; Federico, M.B.; Bertolin, A.P.; Gottifredi, V. Rad51 recombinase prevents Mre11 nuclease-dependent degradation and excessive PrimPol-mediated elongation of nascent DNA after UV irradiation. Proc. Natl. Acad. Sci. USA 2015, 112, E6624–E6633. [Google Scholar] [CrossRef] [PubMed]
  12. Mazouzi, A.; Velimezi, G.; Loizou, J.I. DNA replication stress: Causes, resolution and disease. Exp. Cell Res. 2014, 329, 85–93. [Google Scholar] [CrossRef] [PubMed]
  13. Jekimovs, C.; Bolderson, E.; Suraweera, A.; Adams, M.; O’Byrne, K.J.; Richard, D.J. Chemotherapeutic compounds targeting the DNA double-strand break repair pathways: The good, the bad, and the promising. Front. Oncol. 2014, 4, 86. [Google Scholar] [CrossRef] [PubMed]
  14. Beranek, D.T. Distribution of methyl and ethyl adducts following alkylation with monofunctional alkylating agents. Mutat. Res. 1990, 231, 11–30. [Google Scholar] [CrossRef]
  15. Kondo, N.; Takahashi, A.; Mori, E.; Noda, T.; Su, X.; Ohnishi, K.; McKinnon, P.J.; Sakaki, T.; Nakase, H.; Ono, K.; et al. DNA ligase IV is a potential molecular target in ACNU sensitivity. Cancer Sci. 2010, 101, 1881–1885. [Google Scholar] [CrossRef] [PubMed]
  16. Brookes, P.; Lawley, P.D. The reaction of mono- and di-functional alkylating agents with nucleic acids. Biochem. J. 1961, 80, 496–503. [Google Scholar] [CrossRef] [PubMed]
  17. Lawley, P.D.; Brookes, P. The action of alkylating agents on deoxyribonucleic acid in relation to biological effects of the alkylating agents. Exp. Cell Res. 1963, 24 (Suppl. S9), 512–520. [Google Scholar] [CrossRef]
  18. Noll, D.M.; Mason, T.M.; Miller, P.S. Formation and repair of interstrand cross-links in DNA. Chem. Rev. 2006, 106, 277–301. [Google Scholar] [CrossRef] [PubMed]
  19. Schärer, O.D. DNA Interstrand Crosslinks: Natural and Drug-Induced DNA Adducts that Induce Unique Cellular Responses. ChemBioChem 2005, 6, 27–32. [Google Scholar] [CrossRef] [PubMed]
  20. Lawley, P.D.; Phillips, D.H. DNA adducts from chemotherapeutic agents. Mutat. Res. 1996, 355, 13–40. [Google Scholar] [CrossRef]
  21. Bhuyan, B.K.; Scheidt, L.G.; Fraser, T.J. Cell cycle phase specificity of antitumor agents. Cancer Res. 1972, 32, 398–407. [Google Scholar] [PubMed]
  22. Glover, T.W.; Arlt, M.F.; Casper, A.M.; Durkin, S.G. Mechanisms of common fragile site instability. Hum. Mol. Genet. 2005, 14, R197–R205. [Google Scholar] [CrossRef] [PubMed]
  23. 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. [Google Scholar] [CrossRef] [PubMed]
  24. Hsiang, Y.H.; Lihou, M.G.; Liu, L.F. Arrest of replication forks by drug-stabilized topoisomerase I-DNA cleavable complexes as a mechanism of cell killing by camptothecin. Cancer Res. 1989, 49, 5077–5082. [Google Scholar] [PubMed]
  25. Deweese, J.E.; Osheroff, N. The DNA cleavage reaction of topoisomerase II: Wolf in sheep’s clothing. Nucleic Acids Res. 2009, 37, 738–748. [Google Scholar] [CrossRef] [PubMed]
  26. Helleday, T.; Petermann, E.; Lundin, C.; Hodgson, B.; Sharma, R.A. DNA repair pathways as targets for cancer therapy. Nat. Rev. Cancer 2008, 8, 193–204. [Google Scholar] [CrossRef] [PubMed]
  27. Krokan, H.E.; Bjørås, M. Base Excision Repair. Cold Spring Harb. Perspect. Biol. 2013, 5, a012583. [Google Scholar] [CrossRef] [PubMed]
  28. Gillet, L.C.J.; Schärer, O.D. Molecular mechanisms of mammalian global genome nucleotide excision repair. Chem. Rev. 2006, 106, 253–276. [Google Scholar] [CrossRef] [PubMed]
  29. Caldecott, K.W. Protein ADP-ribosylation and the cellular response to DNA strand breaks. DNA Repair 2014, 19, 108–113. [Google Scholar] [CrossRef] [PubMed]
  30. Heyer, W.-D.; Ehmsen, K.T.; Liu, J. Regulation of homologous recombination in eukaryotes. Annu. Rev. Genet. 2010, 44, 113–139. [Google Scholar] [CrossRef] [PubMed]
  31. Davis, A.J.; Chen, D.J. DNA double strand break repair via non-homologous end-joining. Transl. Cancer Res. 2013, 2, 130–143. [Google Scholar] [PubMed]
  32. Bi, X. Mechanism of DNA damage tolerance. World J. Biol. Chem. 2015, 6, 48–56. [Google Scholar] [CrossRef] [PubMed]
  33. Aguilera, A.; Gómez-González, B. Genome instability: A mechanistic view of its causes and consequences. Nat. Rev. Genet. 2008, 9, 204–217. [Google Scholar] [CrossRef] [PubMed]
  34. Chang, D.J.; Cimprich, K.A. DNA damage tolerance: When it’s OK to make mistakes. Nat. Chem. Biol. 2009, 5, 82–90. [Google Scholar] [CrossRef] [PubMed]
  35. Ghosal, G.; Chen, J. DNA damage tolerance: A double-edged sword guarding the genome. Transl. Cancer Res. 2013, 2, 107–129. [Google Scholar] [PubMed]
  36. Saugar, I.; Ortiz-Bazán, M.Á.; Tercero, J.A. Tolerating DNA damage during eukaryotic chromosome replication. Exp. Cell Res. 2014, 329, 170–177. [Google Scholar] [CrossRef] [PubMed]
  37. Deans, A.J.; West, S.C. DNA interstrand crosslink repair and cancer. Nat. Rev. Cancer 2011, 11, 467–480. [Google Scholar] [CrossRef] [PubMed]
  38. Longerich, S.; Li, J.; Xiong, Y.; Sung, P.; Kupfer, G.M. Stress and DNA repair biology of the Fanconi anemia pathway. Blood 2014, 124, 2812–2819. [Google Scholar] [CrossRef] [PubMed]
  39. Gaillard, H.; García-Muse, T.; Aguilera, A. Replication stress and cancer. Nat. Rev. Cancer 2015, 15, 276–289. [Google Scholar] [CrossRef] [PubMed]
  40. Mamrak, N.E.; Shimamura, A.; Howlett, N.G. Recent discoveries in the molecular pathogenesis of the inherited bone marrow failure syndrome Fanconi anemia. Blood Rev. 2016. [Google Scholar] [CrossRef] [PubMed]
  41. Kennedy, R.D.; D’Andrea, A.D. The Fanconi Anemia/BRCA pathway: New faces in the crowd. Genes Dev. 2005, 19, 2925–2940. [Google Scholar] [CrossRef] [PubMed]
  42. Thompson, L.H.; Hinz, J.M. Cellular and molecular consequences of defective Fanconi anemia proteins in replication-coupled DNA repair: Mechanistic insights. Mutat. Res. 2009, 668, 54–72. [Google Scholar] [CrossRef] [PubMed]
  43. Branzei, D. Ubiquitin family modifications and template switching. FEBS Lett. 2011, 585, 2810–2817. [Google Scholar] [CrossRef] [PubMed]
  44. Xu, X.; Blackwell, S.; Lin, A.; Li, F.; Qin, Z.; Xiao, W. Error-free DNA-damage tolerance in Saccharomyces cerevisiae. Mutat. Res. Rev. Mutat. Res. 2015, 764, 43–50. [Google Scholar] [CrossRef] [PubMed]
  45. Ge, X.Q.; Jackson, D.A.; Blow, J.J. Dormant origins licensed by excess Mcm2–7 are required for human cells to survive replicative stress. Genes Dev. 2007, 21, 3331–3341. [Google Scholar] [CrossRef] [PubMed]
  46. Lopes, M.; Foiani, M.; Sogo, J.M. Multiple mechanisms control chromosome integrity after replication fork uncoupling and restart at irreparable UV lesions. Mol. Cell 2006, 21, 15–27. [Google Scholar] [CrossRef] [PubMed]
  47. Woodward, A.M.; Göhler, T.; Luciani, M.G.; Oehlmann, M.; Ge, X.; Gartner, A.; Jackson, D.A.; Blow, J.J. Excess Mcm2–7 license dormant origins of replication that can be used under conditions of replicative stress. J. Cell Biol. 2006, 173, 673–683. [Google Scholar] [CrossRef] [PubMed]
  48. Elvers, I.; Johansson, F.; Groth, P.; Erixon, K.; Helleday, T. UV stalled replication forks restart by re-priming in human fibroblasts. Nucleic Acids Res. 2011, 39, 7049–7057. [Google Scholar] [CrossRef] [PubMed]
  49. McIntosh, D.; Blow, J.J. Dormant origins, the licensing checkpoint, and the response to replicative stresses. Cold Spring Harb. Perspect. Biol. 2012, 4, a012955. [Google Scholar] [CrossRef] [PubMed]
  50. De Piccoli, G.; Katou, Y.; Itoh, T.; Nakato, R.; Shirahige, K.; Labib, K. Replisome stability at defective DNA replication forks is independent of S phase checkpoint kinases. Mol. Cell 2012, 45, 696–704. [Google Scholar] [CrossRef] [PubMed]
  51. Tercero, J.A.; Diffley, J.F.X. Regulation of DNA replication fork progression through damaged DNA by the Mec1/Rad53 checkpoint. Nature 2001, 412, 553–557. [Google Scholar] [CrossRef] [PubMed]
  52. Lopes, M.; Cotta-Ramusino, C.; Pellicioli, A.; Liberi, G.; Plevani, P.; Muzi-Falconi, M.; Newlon, C.S.; Foiani, M. The DNA replication checkpoint response stabilizes stalled replication forks. Nature 2001, 412, 557–561. [Google Scholar] [CrossRef] [PubMed]
  53. Cobb, J.A.; Bjergbaek, L.; Shimada, K.; Frei, C.; Gasser, S.M. DNA polymerase stabilization at stalled replication forks requires Mec1 and the RecQ helicase Sgs1. EMBO J. 2003, 22, 4325–4336. [Google Scholar] [CrossRef] [PubMed]
  54. 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. [Google Scholar] [CrossRef] [PubMed]
  55. Hanada, K.; Budzowska, M.; Davies, S.L.; van Drunen, E.; Onizawa, H.; Beverloo, H.B.; Maas, A.; Essers, J.; Hickson, I.D.; Kanaar, R. The structure-specific endonuclease Mus81 contributes to replication restart by generating double-strand DNA breaks. Nat. Struct. Mol. Biol. 2007, 14, 1096–1104. [Google Scholar] [CrossRef] [PubMed]
  56. Forment, J.V.; Blasius, M.; Guerini, I.; Jackson, S.P. Structure-specific DNA endonuclease Mus81/Eme1 generates DNA damage caused by Chk1 inactivation. PLoS ONE 2011, 6, e23517. [Google Scholar] [CrossRef] [PubMed]
  57. 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. [Google Scholar] [CrossRef] [PubMed]
  58. Pacek, M.; Walter, J.C. A requirement for MCM7 and Cdc45 in chromosome unwinding during eukaryotic DNA replication. EMBO J. 2004, 23, 3667–3676. [Google Scholar] [CrossRef] [PubMed]
  59. Byun, T.S.; Pacek, M.; Yee, M.; 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. [Google Scholar] [CrossRef] [PubMed]
  60. Zou, L.; Elledge, S.J. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 2003, 300, 1542–1548. [Google Scholar] [CrossRef] [PubMed]
  61. MacDougall, C.A.; Byun, T.S.; Van, C.; Yee, M.; Cimprich, K.A. The structural determinants of checkpoint activation. Genes Dev. 2007, 21, 898–903. [Google Scholar] [CrossRef] [PubMed]
  62. Nam, E.A.; Cortez, D. ATR signalling: More than meeting at the fork. Biochem. J. 2011, 436, 527–536. [Google Scholar] [CrossRef] [PubMed]
  63. Maréchal, A.; Zou, L. DNA damage sensing by the ATM and ATR kinases. Cold Spring Harb. Perspect. Biol. 2013, 5, a012716. [Google Scholar] [CrossRef] [PubMed]
  64. Lucca, C.; Vanoli, F.; Cotta-Ramusino, C.; Pellicioli, A.; Liberi, G.; Haber, J.; Foiani, M. Checkpoint-mediated control of replisome-fork association and signalling in response to replication pausing. Oncogene 2004, 23, 1206–1213. [Google Scholar] [CrossRef] [PubMed]
  65. Petermann, E.; Orta, M.L.; Issaeva, N.; Schultz, N.; Helleday, T. Hydroxyurea-stalled replication forks become progressively inactivated and require two different RAD51-mediated pathways for restart and repair. Mol. Cell 2010, 37, 492–502. [Google Scholar] [CrossRef] [PubMed]
  66. Labib, K.; De Piccoli, G. Surviving chromosome replication: The many roles of the S-phase checkpoint pathway. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2011, 366, 3554–3561. [Google Scholar] [CrossRef] [PubMed]
  67. Ozeri-Galai, E.; Schwartz, M.; Rahat, A.; Kerem, B. Interplay between ATM and ATR in the regulation of common fragile site stability. Oncogene 2008, 27, 2109–2117. [Google Scholar] [CrossRef] [PubMed]
  68. Ciccia, A.; Elledge, S.J. The DNA damage response: Making it safe to play with knives. Mol. Cell 2010, 40, 179–204. [Google Scholar] [CrossRef] [PubMed]
  69. Ammazzalorso, F.; Pirzio, L.M.; Bignami, M.; Franchitto, A.; Pichierri, P. ATR and ATM differently regulate WRN to prevent DSBs at stalled replication forks and promote replication fork recovery. EMBO J. 2010, 29, 3156–3169. [Google Scholar] [CrossRef] [PubMed]
  70. Bachrati, C.Z.; Hickson, I.D. RecQ helicases: Suppressors of tumorigenesis and premature aging. Biochem. J. 2003, 374, 577–606. [Google Scholar] [CrossRef] [PubMed]
  71. Hills, S.A.; Diffley, J.F.X. DNA replication and oncogene-induced replicative stress. Curr. Biol. 2014, 24, R435–R444. [Google Scholar] [CrossRef] [PubMed]
  72. Macheret, M.; Halazonetis, T.D. DNA replication stress as a hallmark of cancer. Annu. Rev. Pathol. 2015, 10, 425–448. [Google Scholar] [CrossRef] [PubMed]
  73. Murga, M.; Campaner, S.; Lopez-Contreras, A.J.; Toledo, L.I.; Soria, R.; Montaña, M.F.; D’Artista, L.; Schleker, T.; Guerra, C.; Garcia, E.; et al. Exploiting oncogene-induced replicative stress for the selective killing of Myc-driven tumors. Nat. Struct. Mol. Biol. 2011, 18, 1331–1335. [Google Scholar] [CrossRef] [PubMed]
  74. Marusyk, A.; Wheeler, L.J.; Mathews, C.K.; DeGregori, J. p53 mediates senescence-like arrest induced by chronic replicational stress. Mol. Cell. Biol. 2007, 27, 5336–5351. [Google Scholar] [CrossRef] [PubMed]
  75. Bai, G.; Smolka, M.B.; Schimenti, J.C. Chronic DNA Replication Stress Reduces Replicative Lifespan of Cells by TRP53-Dependent, microRNA-Assisted MCM2–7 Downregulation. PLoS Genet. 2016, 12, e1005787. [Google Scholar] [CrossRef] [PubMed]
  76. Bartkova, J.; Horejsí, Z.; Koed, K.; Krämer, A.; Tort, F.; Zieger, K.; Guldberg, P.; Sehested, M.; Nesland, J.M.; Lukas, C.; et al. DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 2005, 434, 864–870. [Google Scholar] [CrossRef] [PubMed]
  77. O’Driscoll, M.; Ruiz-Perez, V.L.; Woods, C.G.; Jeggo, P.A.; Goodship, J.A. A splicing mutation affecting expression of ataxia-telangiectasia and Rad3-related protein (ATR) results in Seckel syndrome. Nat. Genet. 2003, 33, 497–501. [Google Scholar] [CrossRef] [PubMed]
  78. McKinnon, P.J. ATM and ataxia telangiectasia. EMBO Rep. 2004, 5, 772–776. [Google Scholar] [CrossRef] [PubMed]
  79. DiGiovanna, J.J.; Kraemer, K.H. Shining a light on xeroderma pigmentosum. J. Investig. Dermatol. 2012, 132, 785–796. [Google Scholar] [CrossRef] [PubMed]
  80. Callén, E.; Surrallés, J. Telomere dysfunction in genome instability syndromes. Mutat. Res. 2004, 567, 85–104. [Google Scholar] [CrossRef] [PubMed]
  81. Lauper, J.M.; Krause, A.; Vaughan, T.L.; Monnat, R.J. Spectrum and risk of neoplasia in Werner syndrome: A systematic review. PLoS ONE 2013, 8, e59709. [Google Scholar] [CrossRef] [PubMed]
  82. Bernstein, K.A.; Gangloff, S.; Rothstein, R. The RecQ DNA helicases in DNA repair. Annu. Rev. Genet. 2010, 44, 393–417. [Google Scholar] [CrossRef] [PubMed]
  83. Kim, H.; D’Andrea, A.D. Regulation of DNA cross-link repair by the Fanconi anemia/BRCA pathway. Genes Dev. 2012, 26, 1393–1408. [Google Scholar] [CrossRef] [PubMed]
  84. Joenje, H.; Patel, K.J. The emerging genetic and molecular basis of Fanconi anaemia. Nat. Rev. Genet. 2001, 2, 446–457. [Google Scholar] [CrossRef] [PubMed]
  85. Larizza, L.; Roversi, G.; Volpi, L. Rothmund-Thomson syndrome. Orphanet J. Rare Dis. 2010, 5, 2. [Google Scholar] [CrossRef] [PubMed]
  86. Lu, H.; Shamanna, R.A.; Keijzers, G.; Anand, R.; Rasmussen, L.J.; Cejka, P.; Croteau, D.L.; Bohr, V.A. RECQL4 Promotes DNA End Resection in Repair of DNA Double-Strand Breaks. Cell Rep. 2016, 16, 161–173. [Google Scholar] [CrossRef] [PubMed]
  87. Rosenberg, B.; Vancamp, L.; Krigas, T. Inhibition of cell division in Escherichia coli by electrolysis products from a platinum electrode. Nature 1965, 205, 698–699. [Google Scholar] [CrossRef] [PubMed]
  88. Todd, R.C.; Lippard, S.J. Inhibition of transcription by platinum antitumor compounds. Metallomics 2009, 1, 280–291. [Google Scholar] [CrossRef] [PubMed]
  89. Zamble, D.B.; Lippard, S.J. Cisplatin and DNA repair in cancer chemotherapy. Trends Biochem. Sci. 1995, 20, 435–439. [Google Scholar] [CrossRef]
  90. Available online: http://www.rcsb.org/pdb/explore/explore.do?structureId=3CO3 (accessed on 23 January 2017).
  91. Fichtinger-Schepman, A.M.; van der Veer, J.L.; den Hartog, J.H.; Lohman, P.H.; Reedijk, J. Adducts of the antitumor drug cis-diamminedichloroplatinum(II) with DNA: Formation, identification, and quantitation. Biochemistry 1985, 24, 707–713. [Google Scholar] [CrossRef] [PubMed]
  92. Harder, H.C.; Rosenberg, B. Inhibitory effects of anti-tumor platinum compounds on DNA, RNA and protein syntheses in mammalian cells in virtro. Int. J. Cancer 1970, 6, 207–216. [Google Scholar] [CrossRef] [PubMed]
  93. Eastman, A. Reevaluation of interaction of cis-dichloro(ethylenediamine)platinum(II) with DNA. Biochemistry 1986, 25, 3912–3915. [Google Scholar] [CrossRef] [PubMed]
  94. Sherman, S.E.; Gibson, D.; Wang, A.H.; Lippard, S.J. X-ray structure of the major adduct of the anticancer drug cisplatin with DNA: cis-[Pt(NH3)2(d(pGpG))]. Science 1985, 230, 412–417. [Google Scholar] [CrossRef] [PubMed]
  95. Eastman, A. Separation and characterization of products resulting from the reaction of cis-diamminedichloroplatinum (II) with deoxyribonucleosides. Biochemistry 1982, 21, 6732–6736. [Google Scholar] [CrossRef] [PubMed]
  96. Desoize, B. Cancer and metals and metal compounds: Part I—Carcinogenesis. Crit. Rev. Oncol. Hematol. 2002, 42, 1–3. [Google Scholar] [CrossRef]
  97. Köberle, B.; Masters, J.R.; Hartley, J.A.; Wood, R.D. Defective repair of cisplatin-induced DNA damage caused by reduced XPA protein in testicular germ cell tumours. Curr. Biol. 1999, 9, 273–276. [Google Scholar] [CrossRef]
  98. Borst, P.; Rottenberg, S.; Jonkers, J. How do real tumors become resistant to cisplatin? Cell Cycle 2008, 7, 1353–1359. [Google Scholar] [CrossRef] [PubMed]
  99. Sedletska, Y.; Fourrier, L.; Malinge, J.-M. Modulation of MutS ATP-dependent functional activities by DNA containing a cisplatin compound lesion (base damage and mismatch). J. Mol. Biol. 2007, 369, 27–40. [Google Scholar] [CrossRef] [PubMed]
  100. Brown, R.; Clugston, C.; Burns, P.; Edlin, A.; Vasey, P.; Vojtĕsek, B.; Kaye, S.B. Increased accumulation of p53 protein in cisplatin-resistant ovarian cell lines. Int. J. Cancer 1993, 55, 678–684. [Google Scholar] [CrossRef] [PubMed]
  101. Damsma, G.E.; Alt, A.; Brueckner, F.; Carell, T.; Cramer, P. Mechanism of transcriptional stalling at cisplatin-damaged DNA. Nat. Struct. Mol. Biol. 2007, 14, 1127–1133. [Google Scholar] [CrossRef] [PubMed]
  102. Shimodaira, H.; Yoshioka-Yamashita, A.; Kolodner, R.D.; Wang, J.Y.J. Interaction of mismatch repair protein PMS2 and the p53-related transcription factor p73 in apoptosis response to cisplatin. Proc. Natl. Acad. Sci. USA 2003, 100, 2420–2425. [Google Scholar] [CrossRef] [PubMed]
  103. Aebi, S.; Kurdi-Haidar, B.; Gordon, R.; Cenni, B.; Zheng, H.; Fink, D.; Christen, R.D.; Boland, C.R.; Koi, M.; Fishel, R.; et al. Loss of DNA mismatch repair in acquired resistance to cisplatin. Cancer Res. 1996, 56, 3087–3090. [Google Scholar] [PubMed]
  104. Alt, A.; Lammens, K.; Chiocchini, C.; Lammens, A.; Pieck, J.C.; Kuch, D.; Hopfner, K.-P.; Carell, T. Bypass of DNA lesions generated during anticancer treatment with cisplatin by DNA polymerase eta. Science 2007, 318, 967–970. [Google Scholar] [CrossRef] [PubMed]
  105. Brozovic, A.; Ambriović-Ristov, A.; Osmak, M. The relationship between cisplatin-induced reactive oxygen species, glutathione, and BCL-2 and resistance to cisplatin. Crit. Rev. Toxicol. 2010, 40, 347–359. [Google Scholar] [CrossRef] [PubMed]
  106. Splettstoesser, F.; Florea, A.-M.; Büsselberg, D. IP3 receptor antagonist, 2-APB, attenuates cisplatin induced Ca2+-influx in HeLa-S3 cells and prevents activation of calpain and induction of apoptosis. Br. J. Pharmacol. 2007, 151, 1176–1186. [Google Scholar] [CrossRef] [PubMed]
  107. Shamimi-Noori, S.; Yeow, W.-S.; Ziauddin, M.F.; Xin, H.; Tran, T.L.N.; Xie, J.; Loehfelm, A.; Patel, P.; Yang, J.; Schrump, D.S.; et al. Cisplatin enhances the antitumor effect of tumor necrosis factor-related apoptosis-inducing ligand gene therapy via recruitment of the mitochondria-dependent death signaling pathway. Cancer Gene Ther. 2008, 15, 356–370. [Google Scholar] [CrossRef] [PubMed]
  108. Qian, W.; Nishikawa, M.; Haque, A.M.; Hirose, M.; Mashimo, M.; Sato, E.; Inoue, M. Mitochondrial density determines the cellular sensitivity to cisplatin-induced cell death. Am. J. Physiol. Cell Physiol. 2005, 289, C1466–C1475. [Google Scholar] [CrossRef] [PubMed]
  109. Wetzel, C.C.; Berberich, S.J. p53 binds to cisplatin-damaged DNA. Biochim. Biophys. Acta 2001, 1517, 392–397. [Google Scholar] [CrossRef]
  110. Kutuk, O.; Arisan, E.D.; Tezil, T.; Shoshan, M.C.; Basaga, H. Cisplatin overcomes Bcl-2-mediated resistance to apoptosis via preferential engagement of Bak: Critical role of Noxa-mediated lipid peroxidation. Carcinogenesis 2009, 30, 1517–1527. [Google Scholar] [CrossRef] [PubMed]
  111. Kim, H.-S.; Hwang, J.-T.; Yun, H.; Chi, S.-G.; Lee, S.-J.; Kang, I.; Yoon, K.-S.; Choe, W.-J.; Kim, S.-S.; Ha, J. Inhibition of AMP-activated protein kinase sensitizes cancer cells to cisplatin-induced apoptosis via hyper-induction of p53. J. Biol. Chem. 2008, 283, 3731–3742. [Google Scholar] [CrossRef] [PubMed]
  112. Yang, C.; Kaushal, V.; Haun, R.S.; Seth, R.; Shah, S.V.; Kaushal, G.P. Transcriptional activation of caspase-6 and -7 genes by cisplatin-induced p53 and its functional significance in cisplatin nephrotoxicity. Cell Death Differ. 2008, 15, 530–544. [Google Scholar] [CrossRef] [PubMed]
  113. Jiang, M.; Wei, Q.; Wang, J.; Du, Q.; Yu, J.; Zhang, L.; Dong, Z. Regulation of PUMA-alpha by p53 in cisplatin-induced renal cell apoptosis. Oncogene 2006, 25, 4056–4066. [Google Scholar] [CrossRef] [PubMed]
  114. Righetti, S.C.; Della Torre, G.; Pilotti, S.; Ménard, S.; Ottone, F.; Colnaghi, M.I.; Pierotti, M.A.; Lavarino, C.; Cornarotti, M.; Oriana, S.; et al. A comparative study of p53 gene mutations, protein accumulation, and response to cisplatin-based chemotherapy in advanced ovarian carcinoma. Cancer Res. 1996, 56, 689–693. [Google Scholar] [PubMed]
  115. Johnson, C.L.; Lu, D.; Huang, J.; Basu, A. Regulation of p53 stabilization by DNA damage and protein kinase C. Mol. Cancer Ther. 2002, 1, 861–867. [Google Scholar] [PubMed]
  116. Gong, J.G.; Costanzo, A.; Yang, H.Q.; Melino, G.; Kaelin, W.G.; Levrero, M.; Wang, J.Y. The tyrosine kinase c-Abl regulates p73 in apoptotic response to cisplatin-induced DNA damage. Nature 1999, 399, 806–809. [Google Scholar] [PubMed]
  117. Preyer, M.; Shu, C.-W.; Wang, J.Y.J. Delayed activation of Bax by DNA damage in embryonic stem cells with knock-in mutations of the Abl nuclear localization signals. Cell Death Differ. 2007, 14, 1139–1148. [Google Scholar] [CrossRef] [PubMed]
  118. Tsai, K.K.C.; Yuan, Z.-M. c-Abl stabilizes p73 by a phosphorylation-augmented interaction. Cancer Res. 2003, 63, 3418–3424. [Google Scholar] [PubMed]
  119. Levy, D.; Adamovich, Y.; Reuven, N.; Shaul, Y. Yap1 phosphorylation by c-Abl is a critical step in selective activation of proapoptotic genes in response to DNA damage. Mol. Cell 2008, 29, 350–361. [Google Scholar] [CrossRef] [PubMed]
  120. Jones, E.V.; Dickman, M.J.; Whitmarsh, A.J. Regulation of p73-mediated apoptosis by c-Jun N-terminal kinase. Biochem. J. 2007, 405, 617–623. [Google Scholar] [CrossRef] [PubMed]
  121. Hayakawa, J.; Ohmichi, M.; Kurachi, H.; Kanda, Y.; Hisamoto, K.; Nishio, Y.; Adachi, K.; Tasaka, K.; Kanzaki, T.; Murata, Y. Inhibition of BAD phosphorylation either at serine 112 via extracellular signal-regulated protein kinase cascade or at serine 136 via Akt cascade sensitizes human ovarian cancer cells to cisplatin. Cancer Res. 2000, 60, 5988–5994. [Google Scholar] [PubMed]
  122. Isonishi, S.; Andrews, P.A.; Howell, S.B. Increased sensitivity to cis-diamminedichloroplatinum(II) in human ovarian carcinoma cells in response to treatment with 12-O-tetradecanoylphorbol 13-acetate. J. Biol. Chem. 1990, 265, 3623–3627. [Google Scholar] [PubMed]
  123. Basu, A.; Teicher, B.A.; Lazo, J.S. Involvement of protein kinase C in phorbol ester-induced sensitization of HeLa cells to cis-diamminedichloroplatinum(II). J. Biol. Chem. 1990, 265, 8451–8457. [Google Scholar] [PubMed]
  124. Wang, X.; Dhalla, N.S. Modification of beta-adrenoceptor signal transduction pathway by genetic manipulation and heart failure. Mol. Cell. Biochem. 2000, 214, 131–155. [Google Scholar] [CrossRef] [PubMed]
  125. Basu, A.; Tu, H. Activation of ERK during DNA damage-induced apoptosis involves protein kinase Cdelta. Biochem. Biophys. Res. Commun. 2005, 334, 1068–1073. [Google Scholar] [CrossRef] [PubMed]
  126. Nowak, G. Protein kinase C-alpha and ERK1/2 mediate mitochondrial dysfunction, decreases in active Na+ transport, and cisplatin-induced apoptosis in renal cells. J. Biol. Chem. 2002, 277, 43377–43388. [Google Scholar] [CrossRef] [PubMed]
  127. Sánchez-Pérez, I.; Benitah, S.A.; Martínez-Gomariz, M.; Lacal, J.C.; Perona, R. Cell stress and MEKK1-mediated c-Jun activation modulate NFκB activity and cell viability. Mol. Biol. Cell 2002, 13, 2933–2945. [Google Scholar] [CrossRef] [PubMed]
  128. Jones, J.A.; Stroud, R.E.; Kaplan, B.S.; Leone, A.M.; Bavaria, J.E.; Gorman, J.H.; Gorman, R.C.; Ikonomidis, J.S. Differential protein kinase C isoform abundance in ascending aortic aneurysms from patients with bicuspid versus tricuspid aortic valves. Circulation 2007, 116, I144–I149. [Google Scholar] [CrossRef] [PubMed]
  129. Zanke, B.W.; Boudreau, K.; Rubie, E.; Winnett, E.; Tibbles, L.A.; Zon, L.; Kyriakis, J.; Liu, F.F.; Woodgett, J.R. The stress-activated protein kinase pathway mediates cell death following injury induced by cis-platinum, UV irradiation or heat. Curr. Biol. 1996, 6, 606–613. [Google Scholar] [CrossRef]
  130. Hernández Losa, J.; Parada Cobo, C.; Guinea Viniegra, J.; Sánchez-Arevalo Lobo, V.J.; Ramón y Cajal, S.; Sánchez-Prieto, R. Role of the p38 MAPK pathway in cisplatin-based therapy. Oncogene 2003, 22, 3998–4006. [Google Scholar] [CrossRef] [PubMed]
  131. Wang, S.J.; Bourguignon, L.Y.W. Hyaluronan-CD44 promotes phospholipase C-mediated Ca2+ signaling and cisplatin resistance in head and neck cancer. Arch. Otolaryngol. Head Neck Surg. 2006, 132, 19–24. [Google Scholar] [CrossRef] [PubMed]
  132. Speelmans, G.; Staffhorst, R.W.; Versluis, K.; Reedijk, J.; de Kruijff, B. Cisplatin complexes with phosphatidylserine in membranes. Biochemistry 1997, 36, 10545–10550. [Google Scholar] [CrossRef] [PubMed]
  133. Huihui, Z.; Baohuai, W.; Youming, Z.; Kui, W. Calorimetric studies on actin polymerization and a comparison of the effects of cisplatin and transplatin. Thermochim. Acta 1995, 265, 31–38. [Google Scholar] [CrossRef]
  134. Chen, X.; Jiang, Y.; Huang, Z.; Li, D.; Chen, X.; Cao, M.; Meng, Q.; Pang, H.; Sun, L.; Zhao, Y.; et al. miRNA-378 reverses chemoresistance to cisplatin in lung adenocarcinoma cells by targeting secreted clusterin. Sci. Rep. 2016, 6, 19455. [Google Scholar] [CrossRef] [PubMed]
  135. Zhu, H.; Wu, H.; Liu, X.; Evans, B.R.; Medina, D.J.; Liu, C.-G.; Yang, J.-M. Role of MicroRNA miR-27a and miR-451 in the regulation of MDR1/P-glycoprotein expression in human cancer cells. Biochem. Pharmacol. 2008, 76, 582–588. [Google Scholar] [CrossRef] [PubMed]
  136. Vanas, V.; Haigl, B.; Stockhammer, V.; Sutterlüty-Fall, H. MicroRNA-21 Increases Proliferation and Cisplatin Sensitivity of Osteosarcoma-Derived Cells. PLoS ONE 2016, 11, e0161023. [Google Scholar] [CrossRef] [PubMed]
  137. Douple, E.B.; Richmond, R.C. Platinum complexes as radiosensitizers of hypoxic mammalian cells. Br. J. Cancer Suppl. 1978, 3, 98–102. [Google Scholar] [PubMed]
  138. Boeckman, H.J.; Trego, K.S.; Turchi, J.J. Cisplatin sensitizes cancer cells to ionizing radiation via inhibition of nonhomologous end joining. Mol. Cancer Res. 2005, 3, 277–285. [Google Scholar] [CrossRef] [PubMed]
  139. Ishibashi, T.; Lippard, S.J. Telomere loss in cells treated with cisplatin. Proc. Natl. Acad. Sci. USA 1998, 95, 4219–4223. [Google Scholar] [CrossRef] [PubMed]
  140. Jordan, P.; Carmo-Fonseca, M. Cisplatin inhibits synthesis of ribosomal RNA in vivo. Nucleic Acids Res. 1998, 26, 2831–2836. [Google Scholar] [CrossRef] [PubMed]
  141. Tofilon, P.J.; Vines, C.M.; Baker, F.L.; Deen, D.F.; Brock, W.A. cis-Diamminedichloroplatinum(II)-induced sister chromatid exchange: An indicator of sensitivity and heterogeneity in primary human tumor cell cultures. Cancer Res. 1986, 46, 6156–6159. [Google Scholar] [PubMed]
  142. Berndtsson, M.; Hägg, M.; Panaretakis, T.; Havelka, A.M.; Shoshan, M.C.; Linder, S. Acute apoptosis by cisplatin requires induction of reactive oxygen species but is not associated with damage to nuclear DNA. Int. J. Cancer 2007, 120, 175–180. [Google Scholar] [CrossRef] [PubMed]
  143. Rouette, A.; Parent, S.; Girouard, J.; Leblanc, V.; Asselin, E. Cisplatin increases B-cell-lymphoma-2 expression via activation of protein kinase C and Akt2 in endometrial cancer cells. Int. J. Cancer 2012, 130, 1755–1767. [Google Scholar] [CrossRef] [PubMed]
  144. Damia, G.; Filiberti, L.; Vikhanskaya, F.; Carrassa, L.; Taya, Y.; D’incalci, M.; Broggini, M. Cisplatinum and taxol induce different patterns of p53 phosphorylation. Neoplasia 2001, 3, 10–16. [Google Scholar] [CrossRef] [PubMed]
  145. Lützkendorf, J.; Wieduwild, E.; Nerger, K.; Lambrecht, N.; Schmoll, H.-J.; Müller-Tidow, C.; Müller, L.P. Resistance for Genotoxic Damage in Mesenchymal Stromal Cells Is Increased by Hypoxia but Not Generally Dependent on p53-Regulated Cell Cycle Arrest. PLoS ONE 2017, 12, e0169921. [Google Scholar] [CrossRef] [PubMed]
  146. Sorenson, C.M.; Barry, M.A.; Eastman, A. Analysis of events associated with cell cycle arrest at G2 phase and cell death induced by cisplatin. J. Natl. Cancer Inst. 1990, 82, 749–755. [Google Scholar] [CrossRef] [PubMed]
  147. Podratz, J.L.; Knight, A.M.; Ta, L.E.; Staff, N.P.; Gass, J.M.; Genelin, K.; Schlattau, A.; Lathroum, L.; Windebank, A.J. Cisplatin induced mitochondrial DNA damage in dorsal root ganglion neurons. Neurobiol. Dis. 2011, 41, 661–668. [Google Scholar] [CrossRef] [PubMed]
  148. Wagner, J.M.; Karnitz, L.M. Cisplatin-induced DNA damage activates replication checkpoint signaling components that differentially affect tumor cell survival. Mol. Pharmacol. 2009, 76, 208–214. [Google Scholar] [CrossRef] [PubMed]
  149. Bürkle, A.; Chen, G.; Küpper, J.H.; Grube, K.; Zeller, W.J. Increased poly(ADP-ribosyl)ation in intact cells by cisplatin treatment. Carcinogenesis 1993, 14, 559–561. [Google Scholar] [CrossRef] [PubMed]
  150. Jennerwein, M.; Andrews, P.A. Drug accumulation and DNA platination in cells exposed to aquated cisplatin species. Cancer Lett. 1994, 81, 215–220. [Google Scholar] [CrossRef]
  151. Shirazi, F.H.; Molepo, J.M.; Stewart, D.J.; Ng, C.E.; Raaphorst, G.P.; Goel, R. Cytotoxicity, accumulation, and efflux of cisplatin and its metabolites in human ovarian carcinoma cells. Toxicol. Appl. Pharmacol. 1996, 140, 211–218. [Google Scholar] [CrossRef] [PubMed]
  152. Hall, M.D.; Telma, K.A.; Chang, K.-E.; Lee, T.D.; Madigan, J.P.; Lloyd, J.R.; Goldlust, I.S.; Hoeschele, J.D.; Gottesman, M.M. Say no to DMSO: Dimethylsulfoxide inactivates cisplatin, carboplatin, and other platinum complexes. Cancer Res. 2014, 74, 3913–3922. [Google Scholar] [CrossRef] [PubMed]
  153. Available online: http://www.webcitation.org/6mEemW8DM (accessed on 23 November 2016).
  154. Massart, C.; Le Tellier, C.; Gibassier, J.; Leclech, G.; Nicol, M. Modulation by dimethyl sulphoxide of the toxicity induced by cis-diamminedichloroplatinum in cultured thyrocytes. Toxicol. Vitro 1993, 7, 87–94. [Google Scholar] [CrossRef]
  155. Wiltshaw, E.; Subramarian, S.; Alexopoulos, C.; Barker, G.H. Cancer of the ovary: A summary of experience with cis-dichlorodiammineplatinum(II) at the Royal Marsden Hospital. Cancer Treat. Rep. 1979, 63, 1545–1548. [Google Scholar] [PubMed]
  156. Galanski, M. Recent developments in the field of anticancer platinum complexes. Recent Pat. Anticancer Drug Discov. 2006, 1, 285–295. [Google Scholar] [CrossRef] [PubMed]
  157. Lebwohl, D.; Canetta, R. Clinical development of platinum complexes in cancer therapy: An historical perspective and an update. Eur. J. Cancer 1998, 34, 1522–1534. [Google Scholar] [CrossRef]
  158. Baetz, T.; Belch, A.; Couban, S.; Imrie, K.; Yau, J.; Myers, R.; Ding, K.; Paul, N.; Shepherd, L.; Iglesias, J.; et al. Gemcitabine, dexamethasone and cisplatin is an active and non-toxic chemotherapy regimen in relapsed or refractory Hodgkin’s disease: A phase II study by the National Cancer Institute of Canada Clinical Trials Group. Ann. Oncol. 2003, 14, 1762–1767. [Google Scholar] [CrossRef] [PubMed]
  159. Crump, M.; Baetz, T.; Couban, S.; Belch, A.; Marcellus, D.; Howson-Jan, K.; Imrie, K.; Myers, R.; Adams, G.; Ding, K.; et al. Gemcitabine, dexamethasone, and cisplatin in patients with recurrent or refractory aggressive histology B-cell non-Hodgkin lymphoma: A Phase II study by the National Cancer Institute of Canada Clinical Trials Group (NCIC-CTG). Cancer 2004, 101, 1835–1842. [Google Scholar] [CrossRef] [PubMed]
  160. Pearson, A.D.J.; Pinkerton, C.R.; Lewis, I.J.; Imeson, J.; Ellershaw, C.; Machin, D.; European Neuroblastoma Study Group; Children’s Cancer and Leukaemia Group (CCLG formerly United Kingdom Children’s Cancer Study Group). High-dose rapid and standard induction chemotherapy for patients aged over 1 year with stage 4 neuroblastoma: A randomised trial. Lancet Oncol. 2008, 9, 247–256. [Google Scholar] [PubMed]
  161. Reichardt, P. The treatment of uterine sarcomas. Ann. Oncol. 2012, 23 (Suppl. S10), x151–x157. [Google Scholar] [CrossRef] [PubMed]
  162. Dadacaridou, M.; Papanicolaou, X.; Maltesas, D.; Megalakaki, C.; Patos, P.; Panteli, K.; Repousis, P.; Mitsouli-Mentzikof, C. Dexamethasone, cyclophosphamide, etoposide and cisplatin (DCEP) for relapsed or refractory multiple myeloma patients. J. BUON 2007, 12, 41–44. [Google Scholar] [PubMed]
  163. Glover, D.; Glick, J.H.; Weiler, C.; Fox, K.; Guerry, D. WR-2721 and high-dose cisplatin: An active combination in the treatment of metastatic melanoma. J. Clin. Oncol. 1987, 5, 574–578. [Google Scholar] [PubMed]
  164. Berghmans, T.; Paesmans, M.; Lalami, Y.; Louviaux, I.; Luce, S.; Mascaux, C.; Meert, A.P.; Sculier, J.P. Activity of chemotherapy and immunotherapy on malignant mesothelioma: A systematic review of the literature with meta-analysis. Lung Cancer 2002, 38, 111–121. [Google Scholar] [CrossRef]
  165. Hanada, K.; Nishijima, K.; Ogata, H.; Atagi, S.; Kawahara, M. Population pharmacokinetic analysis of cisplatin and its metabolites in cancer patients: Possible misinterpretation of covariates for pharmacokinetic parameters calculated from the concentrations of unchanged cisplatin, ultrafiltered platinum and total platinum. Jpn. J. Clin. Oncol. 2001, 31, 179–184. [Google Scholar] [PubMed]
  166. Daugaard, G.; Abildgaard, U. Cisplatin nephrotoxicity. A review. Cancer Chemother. Pharmacol. 1989, 25, 1–9. [Google Scholar] [CrossRef] [PubMed]
  167. Nagai, N.; Kinoshita, M.; Ogata, H.; Tsujino, D.; Wada, Y.; Someya, K.; Ohno, T.; Masuhara, K.; Tanaka, Y.; Kato, K.; et al. Relationship between pharmacokinetics of unchanged cisplatin and nephrotoxicity after intravenous infusions of cisplatin to cancer patients. Cancer Chemother. Pharmacol. 1996, 39, 131–137. [Google Scholar] [CrossRef] [PubMed]
  168. Kartalou, M.; Essigmann, J.M. Mechanisms of resistance to cisplatin. Mutat. Res. 2001, 478, 23–43. [Google Scholar] [CrossRef]
  169. Olszewski, U.; Hamilton, G. A better platinum-based anticancer drug yet to come? Anticancer Agents Med. Chem. 2010, 10, 293–301. [Google Scholar] [CrossRef] [PubMed]
  170. Dieras, V.; Girre, V.; Guilhaume, M.-N.; Laurence, V.; Mignot, L. Oxaliplatin and ovarian cancer. Bull. Cancer 2006, 93 (Suppl. S1), S35–S39. [Google Scholar] [PubMed]
  171. Ganjavi, H.; Gee, M.; Narendran, A.; Parkinson, N.; Krishnamoorthy, M.; Freedman, M.H.; Malkin, D. Adenovirus-mediated p53 gene therapy in osteosarcoma cell lines: Sensitization to cisplatin and doxorubicin. Cancer Gene Ther. 2006, 13, 415–419. [Google Scholar] [CrossRef] [PubMed]
  172. Michels, J.; Vitale, I.; Senovilla, L.; Enot, D.P.; Garcia, P.; Lissa, D.; Olaussen, K.A.; Brenner, C.; Soria, J.-C.; Castedo, M.; et al. Synergistic interaction between cisplatin and PARP inhibitors in non-small cell lung cancer. Cell Cycle 2013, 12, 877–883. [Google Scholar] [CrossRef] [PubMed]
  173. Balmaña, J.; Tung, N.M.; Isakoff, S.J.; Graña, B.; Ryan, P.D.; Saura, C.; Lowe, E.S.; Frewer, P.; Winer, E.; Baselga, J.; et al. Phase I trial of olaparib in combination with cisplatin for the treatment of patients with advanced breast, ovarian and other solid tumors. Ann. Oncol. 2014, 25, 1656–1663. [Google Scholar] [CrossRef] [PubMed]
  174. Sorenson, C.M.; Eastman, A. Influence of cis-diamminedichloroplatinum(II) on DNA synthesis and cell cycle progression in excision repair proficient and deficient Chinese hamster ovary cells. Cancer Res. 1988, 48, 6703–6707. [Google Scholar] [PubMed]
  175. Vichi, P.; Coin, F.; Renaud, J.P.; Vermeulen, W.; Hoeijmakers, J.H.; Moras, D.; Egly, J.M. Cisplatin- and UV-damaged DNA lure the basal transcription factor TFIID/TBP. EMBO J. 1997, 16, 7444–7456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Cullinane, C.; Mazur, S.J.; Essigmann, J.M.; Phillips, D.R.; Bohr, V.A. Inhibition of RNA polymerase II transcription in human cell extracts by cisplatin DNA damage. Biochemistry 1999, 38, 6204–6212. [Google Scholar] [CrossRef] [PubMed]
  177. Kumar, S.; Kumar, A.; Shah, P.P.; Rai, S.N.; Panguluri, S.K.; Kakar, S.S. MicroRNA signature of cis-platin resistant vs. cis-platin sensitive ovarian cancer cell lines. J. Ovarian Res. 2011, 4, 17. [Google Scholar] [PubMed]
  178. Ciarimboli, G.; Ludwig, T.; Lang, D.; Pavenstädt, H.; Koepsell, H.; Piechota, H.-J.; Haier, J.; Jaehde, U.; Zisowsky, J.; Schlatter, E. Cisplatin nephrotoxicity is critically mediated via the human organic cation transporter 2. Am. J. Pathol. 2005, 167, 1477–1484. [Google Scholar] [CrossRef]
  179. Gressette, M.; Vérillaud, B.; Jimenez-Pailhès, A.-S.; Lelièvre, H.; Lo, K.-W.; Ferrand, F.-R.; Gattolliat, C.-H.; Jacquet-Bescond, A.; Kraus-Berthier, L.; Depil, S.; et al. Treatment of Nasopharyngeal Carcinoma Cells with the Histone-Deacetylase Inhibitor Abexinostat: Cooperative Effects with Cis-platin and Radiotherapy on Patient-Derived Xenografts. PLoS ONE 2014, 9, e91325. [Google Scholar] [CrossRef] [PubMed]
  180. Rout, S.R.; Behera, B.; Maiti, T.K.; Mohapatra, S. Multifunctional magnetic calcium phosphate nanoparticles for targeted platin delivery. Dalton Trans. 2012, 41, 10777–10783. [Google Scholar] [CrossRef] [PubMed]
  181. Kitao, H.; Takata, M. Fanconi anemia: A disorder defective in the DNA damage response. Int. J. Hematol. 2011, 93, 417–424. [Google Scholar] [CrossRef] [PubMed]
  182. Sawant, A.; Kothandapani, A.; Zhitkovich, A.; Sobol, R.W.; Patrick, S.M. Role of mismatch repair proteins in the processing of cisplatin interstrand cross-links. DNA Repair 2015, 35, 126–136. [Google Scholar] [CrossRef] [PubMed]
  183. Cheng, C.H.; Kuchta, R.D. DNA polymerase epsilon: Aphidicolin inhibition and the relationship between polymerase and exonuclease activity. Biochemistry 1993, 32, 8568–8574. [Google Scholar] [CrossRef] [PubMed]
  184. Pedrali-Noy, G.; Spadari, S.; Miller-Faurès, A.; Miller, A.O.; Kruppa, J.; Koch, G. Synchronization of HeLa cell cultures by inhibition of DNA polymerase alpha with aphidicolin. Nucleic Acids Res. 1980, 8, 377–387. [Google Scholar] [CrossRef] [PubMed]
  185. Baranovskiy, A.G.; Babayeva, N.D.; Suwa, Y.; Gu, J.; Pavlov, Y.I.; Tahirov, T.H. Structural basis for inhibition of DNA replication by aphidicolin. Nucleic Acids Res. 2014, 42, 14013–14021. [Google Scholar] [CrossRef] [PubMed]
  186. Spadari, S.; Pedrali-Noy, G.; Falaschi, M.C.; Ciarrocchi, G. Control of DNA replication and cell proliferation in eukaryotes by aphidicolin. Toxicol. Pathol. 1984, 12, 143–148. [Google Scholar] [CrossRef] [PubMed]
  187. Available online: http://www.rcsb.org/pdb/explore/explore.do?structureId=4Q5V) (accessed on 23 January 2017).
  188. Chang, D.J.; Lupardus, P.J.; Cimprich, K.A. Monoubiquitination of proliferating cell nuclear antigen induced by stalled replication requires uncoupling of DNA polymerase and mini-chromosome maintenance helicase activities. J. Biol. Chem. 2006, 281, 32081–32088. [Google Scholar] [CrossRef] [PubMed]
  189. Sutherland, G.R. Chromosomal fragile sites. Genet. Anal. Tech. Appl. 1991, 8, 161–166. [Google Scholar] [CrossRef]
  190. Shiraishi, T.; Druck, T.; Mimori, K.; Flomenberg, J.; Berk, L.; Alder, H.; Miller, W.; Huebner, K.; Croce, C.M. Sequence conservation at human and mouse orthologous common fragile regions, FRA3B/FHIT and Fra14A2/Fhit. Proc. Natl. Acad. Sci. USA 2001, 98, 5722–5727. [Google Scholar] [CrossRef] [PubMed]
  191. Hellman, A.; Zlotorynski, E.; Scherer, S.W.; Cheung, J.; Vincent, J.B.; Smith, D.I.; Trakhtenbrot, L.; Kerem, B. A role for common fragile site induction in amplification of human oncogenes. Cancer Cell 2002, 1, 89–97. [Google Scholar] [CrossRef]
  192. Durkin, S.G.; Ragland, R.L.; Arlt, M.F.; Mulle, J.G.; Warren, S.T.; Glover, T.W. Replication stress induces tumor-like microdeletions in FHIT/FRA3B. Proc. Natl. Acad. Sci. USA 2008, 105, 246–251. [Google Scholar] [CrossRef] [PubMed]
  193. Bristow, R.G.; Hill, R.P. Hypoxia and metabolism. Hypoxia, DNA repair and genetic instability. Nat. Rev. Cancer 2008, 8, 180–192. [Google Scholar] [CrossRef] [PubMed]
  194. MacGregor, J.T.; Schlegel, R.; Wehr, C.M.; Alperin, P.; Ames, B.N. Cytogenetic damage induced by folate deficiency in mice is enhanced by caffeine. Proc. Natl. Acad. Sci. USA 1990, 87, 9962–9965. [Google Scholar] [CrossRef] [PubMed]
  195. Koundrioukoff, S.; Carignon, S.; Técher, H.; Letessier, A.; Brison, O.; Debatisse, M. Stepwise activation of the ATR signaling pathway upon increasing replication stress impacts fragile site integrity. PLoS Genet. 2013, 9, e1003643. [Google Scholar] [CrossRef] [PubMed]
  196. Helmrich, A.; Ballarino, M.; Tora, L. Collisions between replication and transcription complexes cause common fragile site instability at the longest human genes. Mol. Cell 2011, 44, 966–977. [Google Scholar] [CrossRef] [PubMed]
  197. Di Micco, R.; Fumagalli, M.; Cicalese, A.; Piccinin, S.; Gasparini, P.; Luise, C.; Schurra, C.; Garre’, M.; Nuciforo, P.G.; Bensimon, A.; et al. Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 2006, 444, 638–642. [Google Scholar] [CrossRef]
  198. Cha, R.S.; Kleckner, N. ATR homolog Mec1 promotes fork progression, thus averting breaks in replication slow zones. Science 2002, 297, 602–606. [Google Scholar] [CrossRef] [PubMed]
  199. Arlt, M.F.; Mulle, J.G.; Schaibley, V.M.; Ragland, R.L.; Durkin, S.G.; Warren, S.T.; Glover, T.W. Replication stress induces genome-wide copy number changes in human cells that resemble polymorphic and pathogenic variants. Am. J. Hum. Genet. 2009, 84, 339–350. [Google Scholar] [CrossRef] [PubMed]
  200. Hardt, N.; Pedrali-Noy, G.; Focher, F.; Spadari, S. Aphidicolin does not inhibit DNA repair synthesis in ultraviolet-irradiated HeLa cells. A radioautographic study. Biochem. J. 1981, 199, 453–455. [Google Scholar] [CrossRef] [PubMed]
  201. Pedrali-Noy, G.; Belvedere, M.; Crepaldi, T.; Focher, F.; Spadari, S. Inhibition of DNA replication and growth of several human and murine neoplastic cells by aphidicolin without detectable effect upon synthesis of immunoglobulins and HLA antigens. Cancer Res. 1982, 42, 3810–3813. [Google Scholar] [PubMed]
  202. Gera, J.F.; Fady, C.; Gardner, A.; Jacoby, F.J.; Briskin, K.B.; Lichtenstein, A. Inhibition of DNA repair with aphidicolin enhances sensitivity of targets to tumor necrosis factor. J. Immunol. 1993, 151, 3746–3757. [Google Scholar] [PubMed]
  203. Waters, R. Aphidicolin: An inhibitor of DNA repair in human fibroblasts. Carcinogenesis 1981, 2, 795–797. [Google Scholar] [CrossRef] [PubMed]
  204. Wang, F.; Stewart, J.; Price, C.M. Human CST abundance determines recovery from diverse forms of DNA damage and replication stress. Cell Cycle 2014, 13, 3488–3498. [Google Scholar] [CrossRef] [PubMed]
  205. Yeo, J.E.; Lee, E.H.; Hendrickson, E.A.; Sobeck, A. CtIP mediates replication fork recovery in a FANCD2-regulated manner. Hum. Mol. Genet. 2014, 23, 3695–3705. [Google Scholar] [CrossRef] [PubMed]
  206. Chaudhury, I.; Stroik, D.R.; Sobeck, A. FANCD2-controlled chromatin access of the Fanconi-associated nuclease FAN1 is crucial for the recovery of stalled replication forks. Mol. Cell. Biol. 2014, 34, 3939–3954. [Google Scholar] [CrossRef] [PubMed]
  207. Hammond, E.M.; Green, S.L.; Giaccia, A.J. Comparison of hypoxia-induced replication arrest with hydroxyurea and aphidicolin-induced arrest. Mutat. Res. 2003, 532, 205–213. [Google Scholar] [CrossRef] [PubMed]
  208. Borel, F.; Lacroix, F.B.; Margolis, R.L. Prolonged arrest of mammalian cells at the G1/S boundary results in permanent S phase stasis. J. Cell Sci. 2002, 115, 2829–2838. [Google Scholar] [PubMed]
  209. Basile, G.; Leuzzi, G.; Pichierri, P.; Franchitto, A. Checkpoint-dependent and independent roles of the Werner syndrome protein in preserving genome integrity in response to mild replication stress. Nucleic Acids Res. 2014, 42, 12628–12639. [Google Scholar] [CrossRef] [PubMed]
  210. Nguyen, G.H.; Dexheimer, T.S.; Rosenthal, A.S.; Chu, W.K.; Singh, D.K.; Mosedale, G.; Bachrati, C.Z.; Schultz, L.; Sakurai, M.; Savitsky, P.; et al. A small molecule inhibitor of the BLM helicase modulates chromosome stability in human cells. Chem. Biol. 2013, 20, 55–62. [Google Scholar] [CrossRef] [PubMed]
  211. Schmidt, L.; Wiedner, M.; Velimezi, G.; Prochazkova, J.; Owusu, M.; Bauer, S.; Loizou, J.I. ATMIN is required for the ATM-mediated signaling and recruitment of 53BP1 to DNA damage sites upon replication stress. DNA Repair 2014, 24, 122–130. [Google Scholar] [CrossRef] [PubMed]
  212. Fujita, M.; Sasanuma, H.; Yamamoto, K.N.; Harada, H.; Kurosawa, A.; Adachi, N.; Omura, M.; Hiraoka, M.; Takeda, S.; Hirota, K. Interference in DNA replication can cause mitotic chromosomal breakage unassociated with double-strand breaks. PLoS ONE 2013, 8, e60043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Beresova, L.; Vesela, E.; Chamrad, I.; Voller, J.; Yamada, M.; Furst, T.; Lenobel, R.; Chroma, K.; Gursky, J.; Krizova, K.; et al. Role of DNA Repair Factor Xeroderma Pigmentosum Protein Group C in Response to Replication Stress As Revealed by DNA Fragile Site Affinity Chromatography and Quantitative Proteomics. J. Proteome Res. 2016, 15, 4505–4517. [Google Scholar] [CrossRef] [PubMed]
  214. Janson, C.; Nyhan, K.; Murnane, J.P. Replication Stress and Telomere Dysfunction Are Present in Cultured Human Embryonic Stem Cells. Cytogenet. Genome Res. 2015, 146, 251–260. [Google Scholar] [CrossRef] [PubMed]
  215. Miron, K.; Golan-Lev, T.; Dvir, R.; Ben-David, E.; Kerem, B. Oncogenes create a unique landscape of fragile sites. Nat. Commun. 2015, 6, 7094. [Google Scholar] [CrossRef] [PubMed]
  216. Murfuni, I.; De Santis, A.; Federico, M.; Bignami, M.; Pichierri, P.; Franchitto, A. Perturbed replication induced genome wide or at common fragile sites is differently managed in the absence of WRN. Carcinogenesis 2012, 33, 1655–1663. [Google Scholar] [CrossRef] [PubMed]
  217. Wilhelm, T.; Magdalou, I.; Barascu, A.; Técher, H.; Debatisse, M.; Lopez, B.S. Spontaneous slow replication fork progression elicits mitosis alterations in homologous recombination-deficient mammalian cells. Proc. Natl. Acad. Sci. USA 2014, 111, 763–768. [Google Scholar] [CrossRef] [PubMed]
  218. Available online: https://www.google.cz/url?sa=t&rct=j&q=&esrc=s&source=web&cd=3&cad=rja&uact=8&ved=0ahUKEwibvoX5jL_QAhULVSwKHQfLCXwQFggsMAI&url=https%3A%2F%2Fwww.sigmaaldrich.com%2Fcontent%2Fdam%2Fsigma-aldrich%2Fdocs%2FSigma%2FDatasheet%2F6%2Fa0781dat.pdf&usg=AFQjCNEPSqAi (accessed on 23 November 2016).
  219. Sessa, C.; Zucchetti, M.; Davoli, E.; Califano, R.; Cavalli, F.; Frustaci, S.; Gumbrell, L.; Sulkes, A.; Winograd, B.; D’Incalci, M. Phase I and clinical pharmacological evaluation of aphidicolin glycinate. J. Natl. Cancer Inst. 1991, 83, 1160–1164. [Google Scholar] [CrossRef] [PubMed]
  220. Edelson, R.E.; Gorycki, P.D.; MacDonald, T.L. The mechanism of aphidicolin bioinactivation by rat liver in vitro systems. Xenobiotica 1990, 20, 273–287. [Google Scholar] [CrossRef] [PubMed]
  221. Santos, G.B.; Krogh, R.; Magalhaes, L.G.; Andricopulo, A.D.; Pupo, M.T.; Emery, F.S. Semisynthesis of new aphidicolin derivatives with high activity against Trypanosoma cruzi. Bioorg. Med. Chem. Lett. 2016, 26, 1205–1208. [Google Scholar] [CrossRef] [PubMed]
  222. Glover, T.W.; Berger, C.; Coyle, J.; Echo, B. DNA polymerase alpha inhibition by aphidicolin induces gaps and breaks at common fragile sites in human chromosomes. Hum. Genet. 1984, 67, 136–142. [Google Scholar] [CrossRef] [PubMed]
  223. Kurose, A.; Tanaka, T.; Huang, X.; Traganos, F.; Darzynkiewicz, Z. Synchronization in the cell cycle by inhibitors of DNA replication induces histone H2AX phosphorylation: An indication of DNA damage. Cell Prolif. 2006, 39, 231–240. [Google Scholar] [CrossRef] [PubMed]
  224. Trenz, K.; Smith, E.; Smith, S.; Costanzo, V. ATM and ATR promote Mre11 dependent restart of collapsed replication forks and prevent accumulation of DNA breaks. EMBO J. 2006, 25, 1764–1774. [Google Scholar] [CrossRef] [PubMed]
  225. Krakoff, I.H.; Brown, N.C.; Reichard, P. Inhibition of ribonucleoside diphosphate reductase by hydroxyurea. Cancer Res. 1968, 28, 1559–1565. [Google Scholar] [PubMed]
  226. Reichard, P. Interactions between deoxyribonucleotide and DNA synthesis. Annu. Rev. Biochem. 1988, 57, 349–374. [Google Scholar] [CrossRef] [PubMed]
  227. 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. [Google Scholar] [CrossRef] [PubMed]
  228. 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. [Google Scholar] [CrossRef]
  229. Bianchi, V.; Pontis, E.; Reichard, P. Changes of deoxyribonucleoside triphosphate pools induced by hydroxyurea and their relation to DNA synthesis. J. Biol. Chem. 1986, 261, 16037–16042. [Google Scholar] [PubMed]
  230. Skog, S.; Tribukait, B.; Wallström, B.; Eriksson, S. Hydroxyurea-induced cell death as related to cell cycle in mouse and human T-lymphoma cells. Cancer Res. 1987, 47, 6490–6493. [Google Scholar] [PubMed]
  231. Akerblom, L. Azidocytidine is incorporated into RNA of 3T6 mouse fibroblasts. FEBS Lett. 1985, 193, 203–207. [Google Scholar] [CrossRef]
  232. Anglana, M.; Apiou, F.; Bensimon, A.; Debatisse, M. Dynamics of DNA replication in mammalian somatic cells: Nucleotide pool modulates origin choice and interorigin spacing. Cell 2003, 114, 385–394. [Google Scholar] [CrossRef]
  233. Barlow, J.H.; Faryabi, R.B.; Callén, E.; Wong, N.; Malhowski, A.; Chen, H.T.; Gutierrez-Cruz, G.; Sun, H.-W.; McKinnon, P.; Wright, G.; et al. Identification of early replicating fragile sites that contribute to genome instability. Cell 2013, 152, 620–632. [Google Scholar] [CrossRef] [PubMed]
  234. Lönn, U.; Lönn, S. Extensive regions of single-stranded DNA in aphidicolin-treated melanoma cells. Biochemistry 1988, 27, 566–570. [Google Scholar] [CrossRef] [PubMed]
  235. Recolin, B.; Van der Laan, S.; Maiorano, D. Role of replication protein A as sensor in activation of the S-phase checkpoint in Xenopus egg extracts. Nucleic Acids Res. 2012, 40, 3431–3442. [Google Scholar] [CrossRef] [PubMed]
  236. Arlt, M.F.; Ozdemir, A.C.; Birkeland, S.R.; Wilson, T.E.; Glover, T.W. Hydroxyurea induces de novo copy number variants in human cells. Proc. Natl. Acad. Sci. USA 2011, 108, 17360–17365. [Google Scholar] [CrossRef] [PubMed]
  237. 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. [Google Scholar] [CrossRef] [PubMed]
  238. Szikriszt, B.; Póti, Á.; Pipek, O.; Krzystanek, M.; Kanu, N.; Molnár, J.; Ribli, D.; Szeltner, Z.; Tusnády, G.E.; Csabai, I.; et al. A comprehensive survey of the mutagenic impact of common cancer cytotoxics. Genome Biol. 2016, 17, 99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Mistrik, M.; Oplustilova, L.; Lukas, J.; Bartek, J. Low-dose DNA damage and replication stress responses quantified by optimized automated single-cell image analysis. Cell Cycle 2009, 8, 2592–2599. [Google Scholar] [CrossRef] [PubMed]
  240. Ohouo, P.Y.; Bastos de Oliveira, F.M.; Liu, Y.; Ma, C.J.; Smolka, M.B. DNA-repair scaffolds dampen checkpoint signalling by counteracting the adaptor Rad9. Nature 2013, 493, 120–124. [Google Scholar] [CrossRef] [PubMed]
  241. Morafraile, E.C.; Diffley, J.F.X.; Tercero, J.A.; Segurado, M. Checkpoint-dependent RNR induction promotes fork restart after replicative stress. Sci. Rep. 2015, 5, 7886. [Google Scholar] [CrossRef] [PubMed]
  242. Kim, H.-S.; Kim, S.-K.; Hromas, R.; Lee, S.-H. The SET Domain Is Essential for Metnase Functions in Replication Restart and the 5’ End of SS-Overhang Cleavage. PLoS ONE 2015, 10, e0139418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Masuda, T.; Xu, X.; Dimitriadis, E.K.; Lahusen, T.; Deng, C.-X. “DNA Binding Region” of BRCA1 Affects Genetic Stability through modulating the Intra-S-Phase Checkpoint. Int. J. Biol. Sci. 2016, 12, 133–143. [Google Scholar] [CrossRef] [PubMed]
  244. Yarden, R.I.; Metsuyanim, S.; Pickholtz, I.; Shabbeer, S.; Tellio, H.; Papa, M.Z. BRCA1-dependent Chk1 phosphorylation triggers partial chromatin disassociation of phosphorylated Chk1 and facilitates S-phase cell cycle arrest. Int. J. Biochem. Cell Biol. 2012, 44, 1761–1769. [Google Scholar] [CrossRef] [PubMed]
  245. Awate, S.; De Benedetti, A. TLK1B mediated phosphorylation of Rad9 regulates its nuclear/cytoplasmic localization and cell cycle checkpoint. BMC Mol. Biol. 2016, 17, 3. [Google Scholar] [CrossRef] [PubMed]
  246. Ahlskog, J.K.; Larsen, B.D.; Achanta, K.; Sørensen, C.S. ATM/ATR-mediated phosphorylation of PALB2 promotes RAD51 function. EMBO Rep. 2016, 17, 671–681. [Google Scholar] [CrossRef] [PubMed]
  247. Molina, B.; Marchetti, F.; Gómez, L.; Ramos, S.; Torres, L.; Ortiz, R.; Altamirano-Lozano, M.; Carnevale, A.; Frias, S. Hydroxyurea induces chromosomal damage in G2 and enhances the clastogenic effect of mitomycin C in Fanconi anemia cells. Environ. Mol. Mutagen. 2015, 56, 457–467. [Google Scholar] [CrossRef] [PubMed]
  248. Croke, M.; Neumann, M.A.; Grotsky, D.A.; Kreienkamp, R.; Yaddanapudi, S.C.; Gonzalo, S. Differences in 53BP1 and BRCA1 regulation between cycling and non-cycling cells. Cell Cycle 2013, 12, 3629–3639. [Google Scholar] [CrossRef] [PubMed]
  249. Yamada, M.; Watanabe, K.; Mistrik, M.; Vesela, E.; Protivankova, I.; Mailand, N.; Lee, M.; Masai, H.; Lukas, J.; Bartek, J. ATR-Chk1-APC/CCdh1-dependent stabilization of Cdc7-ASK (Dbf4) kinase is required for DNA lesion bypass under replication stress. Genes Dev. 2013, 27, 2459–2472. [Google Scholar] [CrossRef] [PubMed]
  250. Hu, L.; Kim, T.M.; Son, M.Y.; Kim, S.-A.; Holland, C.L.; Tateishi, S.; Kim, D.H.; Yew, P.R.; Montagna, C.; Dumitrache, L.C.; et al. Two replication fork maintenance pathways fuse inverted repeats to rearrange chromosomes. Nature 2013, 501, 569–572. [Google Scholar] [CrossRef] [PubMed]
  251. Lou, T.-F.; Singh, M.; Mackie, A.; Li, W.; Pace, B.S. Hydroxyurea generates nitric oxide in human erythroid cells: Mechanisms for gamma-globin gene activation. Exp. Biol. Med. 2009, 234, 1374–1382. [Google Scholar] [CrossRef] [PubMed]
  252. Vassileva, I.; Yanakieva, I.; Peycheva, M.; Gospodinov, A.; Anachkova, B. The mammalian INO80 chromatin remodeling complex is required for replication stress recovery. Nucleic Acids Res. 2014, 42, 9074–9086. [Google Scholar] [CrossRef] [PubMed]
  253. Park, J.I.; Choi, H.S.; Jeong, J.S.; Han, J.Y.; Kim, I.H. Involvement of p38 kinase in hydroxyurea-induced differentiation of K562 cells. Cell Growth Differ. 2001, 12, 481–486. [Google Scholar] [PubMed]
  254. Barthelemy, J.; Hanenberg, H.; Leffak, M. FANCJ is essential to maintain microsatellite structure genome-wide during replication stress. Nucleic Acids Res. 2016, 44, 6803–6816. [Google Scholar] [CrossRef] [PubMed]
  255. Kunnev, D.; Rusiniak, M.E.; Kudla, A.; Freeland, A.; Cady, G.K.; Pruitt, S.C. DNA damage response and tumorigenesis in Mcm2-deficient mice. Oncogene 2010, 29, 3630–3638. [Google Scholar] [CrossRef] [PubMed]
  256. Da Guarda, C.C.; Santiago, R.P.; Pitanga, T.N.; Santana, S.S.; Zanette, D.L.; Borges, V.M.; Goncalves, M.S. Heme changes HIF-α, eNOS and nitrite production in HUVECs after simvastatin, HU, and ascorbic acid therapies. Microvasc. Res. 2016, 106, 128–136. [Google Scholar] [CrossRef] [PubMed]
  257. Leitch, C.; Osdal, T.; Andresen, V.; Molland, M.; Kristiansen, S.; Nguyen, X.N.; Bruserud, Ø.; Gjertsen, B.T.; McCormack, E. Hydroxyurea synergizes with valproic acid in wild-type p53 acute myeloid leukaemia. Oncotarget 2016, 7, 8105–8118. [Google Scholar] [PubMed]
  258. Liu, K.; Graves, J.D.; Scott, J.D.; Li, R.; Lin, W.-C. Akt switches TopBP1 function from checkpoint activation to transcriptional regulation through phosphoserine binding-mediated oligomerization. Mol. Cell. Biol. 2013, 33, 4685–4700. [Google Scholar] [CrossRef] [PubMed]
  259. Available online: https://www.google.cz/url?sa=t&rct=j&q=&esrc=s&source=web&cd=3&cad=rja&uact=8&ved=0ahUKEwiVt4Lulb_QAhUBGSwKHbcOB_kQFggsMAI&url=https%3A%2F%2Fwww.sigmaaldrich.com%2Fcontent%2Fdam%2Fsigma-aldrich%2Fdocs%2FSigma%2FProduct_Information_Sheet%2F2%2Fh8627pis.pdf& (accessed on 23 January 2017).
  260. Segal, J.B.; Strouse, J.J.; Beach, M.C.; Haywood, C.; Witkop, C.; Park, H.; Wilson, R.F.; Bass, E.B.; Lanzkron, S. Hydroxyurea for the Treatment of Sickle Cell Disease; Evidence Reports/Technology Assessments; Agency for Healthcare Research and Quality (US): Rockville, MD, USA, 2008; pp. 1–95.
  261. Kühr, T.; Burgstaller, S.; Apfelbeck, U.; Linkesch, W.; Seewann, H.; Fridrik, M.; Michlmayr, G.; Krieger, O.; Lutz, D.; Lin, W.; et al. A randomized study comparing interferon (IFNα) plus low-dose cytarabine and interferon plus hydroxyurea (HU) in early chronic-phase chronic myeloid leukemia (CML). Leuk. Res. 2003, 27, 405–411. [Google Scholar] [CrossRef]
  262. Aruch, D.; Mascarenhas, J. Contemporary approach to essential thrombocythemia and polycythemia vera. Curr. Opin. Hematol. 2016, 23, 150–160. [Google Scholar] [CrossRef] [PubMed]
  263. Barbui, T.; Finazzi, M.C.; Finazzi, G. Front-line therapy in polycythemia vera and essential thrombocythemia. Blood Rev. 2012, 26, 205–211. [Google Scholar] [CrossRef] [PubMed]
  264. Benito, J.M.; López, M.; Lozano, S.; Ballesteros, C.; González-Lahoz, J.; Soriano, V. Hydroxyurea exerts an anti-proliferative effect on T cells but has no direct impact on cellular activation. Clin. Exp. Immunol. 2007, 149, 171–177. [Google Scholar] [CrossRef] [PubMed]
  265. Gurberg, J.; Bouganim, N.; Shenouda, G.; Zeitouni, A. A case of recurrent anaplastic meningioma of the skull base with radiologic response to hydroxyurea. J. Neurol. Surg. Rep. 2014, 75, e52–e55. [Google Scholar] [CrossRef] [PubMed]
  266. Kiladjian, J.-J.; Chevret, S.; Dosquet, C.; Chomienne, C.; Rain, J.-D. Treatment of polycythemia vera with hydroxyurea and pipobroman: Final results of a randomized trial initiated in 1980. J. Clin. Oncol. 2011, 29, 3907–3913. [Google Scholar] [CrossRef] [PubMed]
  267. Charache, S.; Barton, F.B.; Moore, R.D.; Terrin, M.L.; Steinberg, M.H.; Dover, G.J.; Ballas, S.K.; McMahon, R.P.; Castro, O.; Orringer, E.P. Hydroxyurea and sickle cell anemia. Clinical utility of a myelosuppressive “switching” agent. The Multicenter Study of Hydroxyurea in Sickle Cell Anemia. Medicine 1996, 75, 300–326. [Google Scholar] [CrossRef] [PubMed]
  268. Steinberg, M.H.; McCarthy, W.F.; Castro, O.; Ballas, S.K.; Armstrong, F.D.; Smith, W.; Ataga, K.; Swerdlow, P.; Kutlar, A.; DeCastro, L.; et al. The risks and benefits of long-term use of hydroxyurea in sickle cell anemia: A 17.5 year follow-up. Am. J. Hematol. 2010, 85, 403–408. [Google Scholar] [CrossRef] [PubMed]
  269. Darzynkiewicz, Z.; Halicka, H.D.; Zhao, H.; Podhorecka, M. Cell synchronization by inhibitors of DNA replication induces replication stress and DNA damage response: Analysis by flow cytometry. Methods Mol. Biol. 2011, 761, 85–96. [Google Scholar] [PubMed]
  270. Fugger, K.; Mistrik, M.; Danielsen, J.R.; Dinant, C.; Falck, J.; Bartek, J.; Lukas, J.; Mailand, N. Human Fbh1 helicase contributes to genome maintenance via pro- and anti-recombinase activities. J. Cell Biol. 2009, 186, 655–663. [Google Scholar] [CrossRef] [PubMed]
  271. Liu, N.; Lim, C.-S. Differential roles of XRCC2 in homologous recombinational repair of stalled replication forks. J. Cell. Biochem. 2005, 95, 942–954. [Google Scholar] [CrossRef] [PubMed]
  272. Brose, R.D.; Shin, G.; McGuinness, M.C.; Schneidereith, T.; Purvis, S.; Dong, G.X.; Keefer, J.; Spencer, F.; Smith, K.D. Activation of the stress proteome as a mechanism for small molecule therapeutics. Hum. Mol. Genet. 2012, 21, 4237–4252. [Google Scholar] [CrossRef] [PubMed]
  273. 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. [Google Scholar] [PubMed]
  274. Wall, M.E.; Wani, M.C.; Cook, C.E.; Palmer, K.H.; McPhail, A.T.; Sim, G.A. Plant Antitumor Agents. I. The Isolation and Structure of Camptothecin, a Novel Alkaloidal Leukemia and Tumor Inhibitor from Camptotheca acuminata1,2. J. Am. Chem. Soc. 1966, 88, 3888–3890. [Google Scholar] [CrossRef]
  275. Gupta, M.; Fujimori, A.; Pommier, Y. Eukaryotic DNA topoisomerases I. Biochim. Biophys. Acta 1995, 1262, 1–14. [Google Scholar] [CrossRef]
  276. Champoux, J.J. Mechanism of the reaction catalyzed by the DNA untwisting enzyme: Attachment of the enzyme to 3′-terminus of the nicked DNA. J. Mol. Biol. 1978, 118, 441–446. [Google Scholar] [CrossRef]
  277. Available online: http://www.rcsb.org/pdb/explore/explore.do?structureId=1T8I (accessed on 23 January 2017).
  278. Stivers, J.T.; Harris, T.K.; Mildvan, A.S. Vaccinia DNA topoisomerase I: Evidence supporting a free rotation mechanism for DNA supercoil relaxation. Biochemistry 1997, 36, 5212–5222. [Google Scholar] [CrossRef] [PubMed]
  279. Koster, D.A.; Palle, K.; Bot, E.S.M.; Bjornsti, M.-A.; Dekker, N.H. Antitumour drugs impede DNA uncoiling by topoisomerase I. Nature 2007, 448, 213–217. [Google Scholar] [CrossRef] [PubMed]
  280. Staker, B.L.; Hjerrild, K.; Feese, M.D.; Behnke, C.A.; Burgin, A.B.; Stewart, L. The mechanism of topoisomerase I poisoning by a camptothecin analog. Proc. Natl. Acad. Sci. USA 2002, 99, 15387–15392. [Google Scholar] [CrossRef] [PubMed]
  281. Regairaz, M.; Zhang, Y.-W.; Fu, H.; Agama, K.K.; Tata, N.; Agrawal, S.; Aladjem, M.I.; Pommier, Y. Mus81-mediated DNA cleavage resolves replication forks stalled by topoisomerase I–DNA complexes. J. Cell Biol. 2011, 195, 739–749. [Google Scholar] [CrossRef] [PubMed]
  282. Palle, K.; Vaziri, C. Rad18 E3 ubiquitin ligase activity mediates Fanconi anemia pathway activation and cell survival following DNA Topoisomerase 1 inhibition. Cell Cycle 2011, 10, 1625–1638. [Google Scholar] [CrossRef] [PubMed]
  283. Pommier, Y. Topoisomerase I inhibitors: Camptothecins and beyond. Nat. Rev. Cancer 2006, 6, 789–802. [Google Scholar] [CrossRef] [PubMed]
  284. Tuduri, S.; Crabbé, L.; Conti, C.; Tourrière, H.; Holtgreve-Grez, H.; Jauch, A.; Pantesco, V.; De Vos, J.; Thomas, A.; Theillet, C.; et al. Topoisomerase I suppresses genomic instability by preventing interference between replication and transcription. Nat. Cell Biol. 2009, 11, 1315–1324. [Google Scholar] [CrossRef] [PubMed]
  285. Tripathi, K.; Mani, C.; Clark, D.W.; Palle, K. Rad18 is required for functional interactions between FANCD2, BRCA2, and Rad51 to repair DNA topoisomerase 1-poisons induced lesions and promote fork recovery. Oncotarget 2016, 7, 12537–12553. [Google Scholar] [PubMed]
  286. Tsao, Y.P.; D’Arpa, P.; Liu, L.F. The involvement of active DNA synthesis in camptothecin-induced G2 arrest: Altered regulation of p34cdc2/cyclin B. Cancer Res. 1992, 52, 1823–1829. [Google Scholar] [PubMed]
  287. Kharbanda, S.; Rubin, E.; Gunji, H.; Hinz, H.; Giovanella, B.; Pantazis, P.; Kufe, D. Camptothecin and its derivatives induce expression of the c-jun protooncogene in human myeloid leukemia cells. Cancer Res. 1991, 51, 6636–6642. [Google Scholar] [PubMed]
  288. Aller, P.; Rius, C.; Mata, F.; Zorrilla, A.; Cabañas, C.; Bellón, T.; Bernabeu, C. Camptothecin induces differentiation and stimulates the expression of differentiation-related genes in U-937 human promonocytic leukemia cells. Cancer Res. 1992, 52, 1245–1251. [Google Scholar] [PubMed]
  289. Clements, M.K.; Jones, C.B.; Cumming, M.; Daoud, S.S. Antiangiogenic potential of camptothecin and topotecan. Cancer Chemother. Pharmacol. 1999, 44, 411–416. [Google Scholar] [CrossRef] [PubMed]
  290. O’Leary, J.J.; Shapiro, R.L.; Ren, C.J.; Chuang, N.; Cohen, H.W.; Potmesil, M. Antiangiogenic effects of camptothecin analogues 9-amino-20(S)-camptothecin, topotecan, and CPT-11 studied in the mouse cornea model. Clin. Cancer Res. 1999, 5, 181–187. [Google Scholar] [PubMed]
  291. Arlt, M.F.; Glover, T.W. Inhibition of topoisomerase I prevents chromosome breakage at common fragile sites. DNA Repair 2010, 9, 678–689. [Google Scholar] [CrossRef] [PubMed]
  292. Horwitz, S.B.; Horwitz, M.S. Effects of camptothecin on the breakage and repair of DNA during the cell cycle. Cancer Res. 1973, 33, 2834–2836. [Google Scholar] [PubMed]
  293. Jayasooriya, R.G.P.T.; Choi, Y.H.; Hyun, J.W.; Kim, G.-Y. Camptothecin sensitizes human hepatoma Hep3B cells to TRAIL-mediated apoptosis via ROS-dependent death receptor 5 upregulation with the involvement of MAPKs. Environ. Toxicol. Pharmacol. 2014, 38, 959–967. [Google Scholar] [CrossRef] [PubMed]
  294. Strumberg, D.; Pilon, A.A.; Smith, M.; Hickey, R.; Malkas, L.; Pommier, Y. Conversion of topoisomerase I cleavage complexes on the leading strand of ribosomal DNA into 5’-phosphorylated DNA double-strand breaks by replication runoff. Mol. Cell. Biol. 2000, 20, 3977–3987. [Google Scholar] [CrossRef] [PubMed]
  295. Priel, E.; Showalter, S.D.; Roberts, M.; Oroszlan, S.; Blair, D.G. The topoisomerase I inhibitor, camptothecin, inhibits equine infectious anemia virus replication in chronically infected CF2Th cells. J. Virol. 1991, 65, 4137–4141. [Google Scholar] [PubMed]
  296. Bruno, S.; Giaretti, W.; Darzynkiewicz, Z. Effect of camptothecin on mitogenic stimulation of human lymphocytes: Involvement of DNA topoisomerase I in cell transition from G0 to G1 phase of the cell cycle and in DNA replication. J. Cell. Physiol. 1992, 151, 478–486. [Google Scholar] [CrossRef] [PubMed]
  297. Squires, S.; Ryan, A.J.; Strutt, H.L.; Johnson, R.T. Hypersensitivity of Cockayne’s syndrome cells to camptothecin is associated with the generation of abnormally high levels of double strand breaks in nascent DNA. Cancer Res. 1993, 53, 2012–2019. [Google Scholar] [PubMed]
  298. Ding, X.; Matsuo, K.; Xu, L.; Yang, J.; Zheng, L. Optimized combinations of bortezomib, camptothecin, and doxorubicin show increased efficacy and reduced toxicity in treating oral cancer. Anticancer Drugs 2015, 26, 547–554. [Google Scholar] [CrossRef] [PubMed]
  299. Zhang, J.; Walter, J.C. Mechanism and regulation of incisions during DNA interstrand cross-link repair. DNA Repair 2014, 19, 135–142. [Google Scholar] [CrossRef] [PubMed]
  300. Ray Chaudhuri, A.; Hashimoto, Y.; Herrador, R.; Neelsen, K.J.; Fachinetti, D.; Bermejo, R.; Cocito, A.; Costanzo, V.; Lopes, M. Topoisomerase I poisoning results in PARP-mediated replication fork reversal. Nat. Struct. Mol. Biol. 2012, 19, 417–423. [Google Scholar] [CrossRef] [PubMed]
  301. Available online: http://www.sigmaaldrich.com/content/dam/sigma-aldrich/docs/Sigma/Product_Information_Sheet/c9911pis.pdf (accessed on 23 November 2016).
  302. Jaxel, C.; Kohn, K.W.; Wani, M.C.; Wall, M.E.; Pommier, Y. Structure-activity study of the actions of camptothecin derivatives on mammalian topoisomerase I: Evidence for a specific receptor site and a relation to antitumor activity. Cancer Res. 1989, 49, 1465–1469. [Google Scholar] [PubMed]
  303. Takagi, K.; Dexheimer, T.S.; Redon, C.; Sordet, O.; Agama, K.; Lavielle, G.; Pierré, A.; Bates, S.E.; Pommier, Y. Novel E-ring camptothecin keto analogues (S38809 and S39625) are stable, potent, and selective topoisomerase I inhibitors without being substrates of drug efflux transporters. Mol. Cancer Ther. 2007, 6, 3229–3238. [Google Scholar] [CrossRef] [PubMed]
  304. Hande, K.R. Etoposide: Four decades of development of a topoisomerase II inhibitor. Eur. J. Cancer 1998, 34, 1514–1521. [Google Scholar] [CrossRef]
  305. Available online: http://www.rcsb.org/pdb/explore/explore.do?structureId=3QX3 (accessed on 23 January 2017).
  306. Liu, L.F.; Rowe, T.C.; Yang, L.; Tewey, K.M.; Chen, G.L. Cleavage of DNA by mammalian DNA topoisomerase II. J. Biol. Chem. 1983, 258, 15365–15370. [Google Scholar] [PubMed]
  307. Gibson, E.G.; King, M.M.; Mercer, S.L.; Deweese, J.E. Two-Mechanism Model for the Interaction of Etoposide Quinone with Topoisomerase IIα. Chem. Res. Toxicol. 2016, 29, 1541–1548. [Google Scholar] [CrossRef] [PubMed]
  308. Wu, C.-C.; Li, T.-K.; Farh, L.; Lin, L.-Y.; Lin, T.-S.; Yu, Y.-J.; Yen, T.-J.; Chiang, C.-W.; Chan, N.-L. Structural basis of type II topoisomerase inhibition by the anticancer drug etoposide. Science 2011, 333, 459–462. [Google Scholar] [CrossRef] [PubMed]
  309. Bender, R.P.; Jablonksy, M.J.; Shadid, M.; Romaine, I.; Dunlap, N.; Anklin, C.; Graves, D.E.; Osheroff, N. Substituents on etoposide that interact with human topoisomerase IIalpha in the binary enzyme-drug complex: Contributions to etoposide binding and activity. Biochemistry 2008, 47, 4501–4509. [Google Scholar] [CrossRef] [PubMed]
  310. Wilstermann, A.M.; Bender, R.P.; Godfrey, M.; Choi, S.; Anklin, C.; Berkowitz, D.B.; Osheroff, N.; Graves, D.E. Topoisomerase II—Drug interaction domains: Identification of substituents on etoposide that interact with the enzyme. Biochemistry 2007, 46, 8217–8225. [Google Scholar] [CrossRef] [PubMed]
  311. Jacob, D.A.; Mercer, S.L.; Osheroff, N.; Deweese, J.E. Etoposide quinone is a redox-dependent topoisomerase II poison. Biochemistry 2011, 50, 5660–5667. [Google Scholar] [CrossRef] [PubMed]
  312. Rogakou, E.P.; Pilch, D.R.; Orr, A.H.; Ivanova, V.S.; Bonner, W.M. DNA double-stranded breaks induce histone H2AX phosphorylation on serine 139. J. Biol. Chem. 1998, 273, 5858–5868. [Google Scholar] [CrossRef] [PubMed]
  313. Terasawa, M.; Shinohara, A.; Shinohara, M. Canonical non-homologous end joining in mitosis induces genome instability and is suppressed by M-phase-specific phosphorylation of XRCC4. PLoS Genet. 2014, 10, e1004563. [Google Scholar] [CrossRef] [PubMed]
  314. Zhao, H.; Rybak, P.; Dobrucki, J.; Traganos, F.; Darzynkiewicz, Z. Relationship of DNA damage signaling to DNA replication following treatment with DNA topoisomerase inhibitors camptothecin/topotecan, mitoxantrone, or etoposide. Cytometry A 2012, 81, 45–51. [Google Scholar] [CrossRef] [PubMed]
  315. Montecucco, A.; Rossi, R.; Ferrari, G.; Scovassi, A.I.; Prosperi, E.; Biamonti, G. Etoposide Induces the Dispersal of DNA Ligase I from Replication Factories. Mol. Biol. Cell 2001, 12, 2109–2118. [Google Scholar] [CrossRef] [PubMed]
  316. Holm, C.; Covey, J.M.; Kerrigan, D.; Pommier, Y. Differential requirement of DNA replication for the cytotoxicity of DNA topoisomerase I and II inhibitors in Chinese hamster DC3F cells. Cancer Res. 1989, 49, 6365–6368. [Google Scholar] [PubMed]
  317. Austin, C.A.; Sng, J.H.; Patel, S.; Fisher, L.M. Novel HeLa topoisomerase II is the II beta isoform: Complete coding sequence and homology with other type II topoisomerases. Biochim. Biophys. Acta 1993, 1172, 283–291. [Google Scholar] [CrossRef]
  318. Niimi, A.; Suka, N.; Harata, M.; Kikuchi, A.; Mizuno, S. Co-localization of chicken DNA topoisomerase IIalpha, but not beta, with sites of DNA replication and possible involvement of a C-terminal region of alpha through its binding to PCNA. Chromosoma 2001, 110, 102–114. [Google Scholar] [CrossRef] [PubMed]
  319. Ju, B.-G.; Lunyak, V.V.; Perissi, V.; Garcia-Bassets, I.; Rose, D.W.; Glass, C.K.; Rosenfeld, M.G. A topoisomerase IIβ-mediated dsDNA break required for regulated transcription. Science 2006, 312, 1798–1802. [Google Scholar] [CrossRef] [PubMed]
  320. Azarova, A.M.; Lyu, Y.L.; Lin, C.-P.; Tsai, Y.-C.; Lau, J.Y.-N.; Wang, J.C.; Liu, L.F. Roles of DNA topoisomerase II isozymes in chemotherapy and secondary malignancies. Proc. Natl. Acad. Sci. USA 2007, 104, 11014–11019. [Google Scholar] [CrossRef] [PubMed]
  321. Nitiss, J.L. DNA topoisomerase II and its growing repertoire of biological functions. Nat. Rev. Cancer 2009, 9, 327–337. [Google Scholar] [CrossRef] [PubMed]
  322. Gupta, R.S.; Bromke, A.; Bryant, D.W.; Gupta, R.; Singh, B.; McCalla, D.R. Etoposide (VP16) and teniposide (VM26): Novel anticancer drugs, strongly mutagenic in mammalian but not prokaryotic test systems. Mutagenesis 1987, 2, 179–186. [Google Scholar] [CrossRef] [PubMed]
  323. Muslimović, A.; Nyström, S.; Gao, Y.; Hammarsten, O. Numerical Analysis of Etoposide Induced DNA Breaks. PLoS ONE 2009, 4, e5859. [Google Scholar] [CrossRef]
  324. Álvarez-Quilón, A.; Serrano-Benítez, A.; Lieberman, J.A.; Quintero, C.; Sánchez-Gutiérrez, D.; Escudero, L.M.; Cortés-Ledesma, F. ATM specifically mediates repair of double-strand breaks with blocked DNA ends. Nat. Commun. 2014, 5, 3347. [Google Scholar] [CrossRef] [PubMed]
  325. Nagano, T.; Nakano, M.; Nakashima, A.; Onishi, K.; Yamao, S.; Enari, M.; Kikkawa, U.; Kamada, S. Identification of cellular senescence-specific genes by comparative transcriptomics. Sci. Rep. 2016, 6, 31758. [Google Scholar] [CrossRef] [PubMed]
  326. Brasacchio, D.; Alsop, A.E.; Noori, T.; Lufti, M.; Iyer, S.; Simpson, K.J.; Bird, P.I.; Kluck, R.M.; Johnstone, R.W.; Trapani, J.A. Epigenetic control of mitochondrial cell death through PACS1-mediated regulation of BAX/BAK oligomerization. Cell Death Differ. 2017. [Google Scholar] [CrossRef] [PubMed]
  327. Martin, R.; Desponds, C.; Eren, R.O.; Quadroni, M.; Thome, M.; Fasel, N. Caspase-mediated cleavage of raptor participates in the inactivation of mTORC1 during cell death. Cell Death Discov. 2016, 2, 16024. [Google Scholar] [CrossRef] [PubMed]
  328. Brekman, A.; Singh, K.E.; Polotskaia, A.; Kundu, N.; Bargonetti, J. A p53-independent role of Mdm2 in estrogen-mediated activation of breast cancer cell proliferation. Breast Cancer Res. 2011, 13, R3. [Google Scholar] [CrossRef] [PubMed]
  329. Soubeyrand, S.; Pope, L.; Haché, R.J.G. Topoisomerase IIα-dependent induction of a persistent DNA damage response in response to transient etoposide exposure. Mol. Oncol. 2010, 4, 38–51. [Google Scholar] [CrossRef] [PubMed]
  330. Velma, V.; Carrero, Z.I.; Allen, C.B.; Hebert, M.D. Coilin levels modulate cell cycle progression and γH2AX levels in etoposide treated U2OS cells. FEBS Lett. 2012, 586, 3404–3409. [Google Scholar] [CrossRef] [PubMed]
  331. Dehennaut, V.; Loison, I.; Dubuissez, M.; Nassour, J.; Abbadie, C.; Leprince, D. DNA double-strand breaks lead to activation of hypermethylated in cancer 1 (HIC1) by SUMOylation to regulate DNA repair. J. Biol. Chem. 2013, 288, 10254–10264. [Google Scholar] [CrossRef] [PubMed]
  332. Paget, S.; Dubuissez, M.; Dehennaut, V.; Nassour, J.; Harmon, B.T.; Spruyt, N.; Loison, I.; Abbadie, C.; Rood, B.R.; Leprince, D. HIC1 (hypermethylated in cancer 1) SUMOylation is dispensable for DNA repair but is essential for the apoptotic DNA damage response (DDR) to irreparable DNA double-strand breaks (DSBs). Oncotarget 2017, 8, 2916–2935. [Google Scholar] [CrossRef] [PubMed]
  333. Sypniewski, D.; Bednarek, I.; Gałka, S.; Loch, T.; Błaszczyk, D.; Sołtysik, D. Cytotoxicity of etoposide in cancer cell lines in vitro after BCL-2 and C-RAF gene silencing with antisense oligonucleotides. Acta Pol. Pharm. 2013, 70, 87–97. [Google Scholar]
  334. Rybak, P.; Hoang, A.; Bujnowicz, L.; Bernas, T.; Berniak, K.; ZarÄ Bski, M.A.; Darzynkiewicz, Z.; Dobrucki, J. Low level phosphorylation of histone H2AX on serine 139 (γH2AX) is not associated with DNA double-strand breaks. Oncotarget 2016, 7, 49574–49587. [Google Scholar] [CrossRef] [PubMed]
  335. Chen, L.; Cui, H.; Fang, J.; Deng, H.; Kuang, P.; Guo, H.; Wang, X.; Zhao, L. Glutamine deprivation plus BPTES alters etoposide- and cisplatin-induced apoptosis in triple negative breast cancer cells. Oncotarget 2016, 7, 54691–54701. [Google Scholar] [CrossRef] [PubMed]
  336. Rodriguez-Lopez, A.M.; Xenaki, D.; Eden, T.O.; Hickman, J.A.; Chresta, C.M. MDM2 mediated nuclear exclusion of p53 attenuates etoposide-induced apoptosis in neuroblastoma cells. Mol. Pharmacol. 2001, 59, 135–143. [Google Scholar]
  337. Litwiniec, A.; Gackowska, L.; Helmin-Basa, A.; Żuryń, A.; Grzanka, A. Low-dose etoposide-treatment induces endoreplication and cell death accompanied by cytoskeletal alterations in A549 cells: Does the response involve senescence? The possible role of vimentin. Cancer Cell Int. 2013, 13, 9. [Google Scholar] [CrossRef] [PubMed]
  338. Akhtar, N.; Talegaonkar, S.; Khar, R.K.; Jaggi, M. A validated stability-indicating LC method for estimation of etoposide in bulk and optimized self-nano emulsifying formulation: Kinetics and stability effects. Saudi Pharm. J. 2013, 21, 103–111. [Google Scholar] [CrossRef] [PubMed]
  339. Available online: http://www.sigmaaldrich.com/content/dam/sigma-aldrich/docs/Sigma/Product_Information_Sheet/e1383pis.pdf (accessed on 23 Novermber 2016).
  340. Wrasidlo, W.; Schröder, U.; Bernt, K.; Hübener, N.; Shabat, D.; Gaedicke, G.; Lode, H. Synthesis, hydrolytic activation and cytotoxicity of etoposide prodrugs. Bioorg Med Chem Lett. 2002, 12, 557–560. [Google Scholar] [CrossRef]
  341. Jokić, M.; Vlašić, I.; Rinneburger, M.; Klümper, N.; Spiro, J.; Vogel, W.; Offermann, A.; Kümpers, C.; Fritz, C.; Schmitt, A.; et al. Ercc1 Deficiency Promotes Tumorigenesis and Increases Cisplatin Sensitivity in a Tp53 Context-Specific Manner. Mol. Cancer Res. 2016, 14, 1110–1123. [Google Scholar] [CrossRef] [PubMed]
  342. Felix, C.A.; Walker, A.H.; Lange, B.J.; Williams, T.M.; Winick, N.J.; Cheung, N.K.; Lovett, B.D.; Nowell, P.C.; Blair, I.A.; Rebbeck, T.R. Association of CYP3A4 genotype with treatment-related leukemia. Proc. Natl. Acad. Sci. USA 1998, 95, 13176–13181. [Google Scholar] [CrossRef] [PubMed]
  343. Blanco, J.G.; Edick, M.J.; Relling, M.V. Etoposide induces chimeric Mll gene fusions. FASEB J. 2004, 18, 173–175. [Google Scholar] [CrossRef] [PubMed]
  344. Thirman, M.J.; Gill, H.J.; Burnett, R.C.; Mbangkollo, D.; McCabe, N.R.; Kobayashi, H.; Ziemin-van der Poel, S.; Kaneko, Y.; Morgan, R.; Sandberg, A.A. Rearrangement of the MLL gene in acute lymphoblastic and acute myeloid leukemias with 11q23 chromosomal translocations. N. Engl. J. Med. 1993, 329, 909–914. [Google Scholar] [CrossRef] [PubMed]
  345. Cerveira, N.; Lisboa, S.; Correia, C.; Bizarro, S.; Santos, J.; Torres, L.; Vieira, J.; Barros-Silva, J.D.; Pereira, D.; Moreira, C.; et al. Genetic and clinical characterization of 45 acute leukemia patients with MLL gene rearrangements from a single institution. Mol. Oncol. 2012, 6, 553–564. [Google Scholar] [CrossRef] [PubMed]
  346. Krivtsov, A.V.; Armstrong, S.A. MLL translocations, histone modifications and leukaemia stem-cell development. Nat. Rev. Cancer 2007, 7, 823–833. [Google Scholar] [CrossRef] [PubMed]
  347. Zhang, L.; Chen, F.; Zhang, Z.; Chen, Y.; Lin, Y.; Wang, J. Design, synthesis and evaluation of the multidrug resistance-reversing activity of pyridine acid esters of podophyllotoxin in human leukemia cells. Bioorg. Med. Chem. Lett. 2016, 26, 4466–4471. [Google Scholar] [CrossRef] [PubMed]
  348. Lee, K.-I.; Su, C.-C.; Yang, C.-Y.; Hung, D.-Z.; Lin, C.-T.; Lu, T.-H.; Liu, S.-H.; Huang, C.-F. Etoposide induces pancreatic β-cells cytotoxicity via the JNK/ERK/GSK-3 signaling-mediated mitochondria-dependent apoptosis pathway. Toxicol. Vitro 2016, 36, 142–152. [Google Scholar] [CrossRef] [PubMed]
  349. Pellegrini, G.G.; Morales, C.C.; Wallace, T.C.; Plotkin, L.I.; Bellido, T. Avenanthramides Prevent Osteoblast and Osteocyte Apoptosis and Induce Osteoclast Apoptosis in Vitro in an Nrf2-Independent Manner. Nutrients 2016, 8, 423. [Google Scholar] [CrossRef] [PubMed]
  350. Papież, M.A.; Krzyściak, W.; Szade, K.; Bukowska-Straková, K.; Kozakowska, M.; Hajduk, K.; Bystrowska, B.; Dulak, J.; Jozkowicz, A. Curcumin enhances the cytogenotoxic effect of etoposide in leukemia cells through induction of reactive oxygen species. Drug Des. Dev. Ther. 2016, 10, 557–570. [Google Scholar] [CrossRef] [PubMed]
  351. Zhang, S.; Lu, C.; Zhang, X.; Li, J.; Jiang, H. Targeted delivery of etoposide to cancer cells by folate-modified nanostructured lipid drug delivery system. Drug Deliv. 2016, 23, 1838–1845. [Google Scholar] [CrossRef] [PubMed]
  352. Lindsay, G.S.; Wallace, H.M. Changes in polyamine catabolism in HL-60 human promyelogenous leukaemic cells in response to etoposide-induced apoptosis. Biochem. J. 1999, 337 Pt 1, 83–87. [Google Scholar] [CrossRef] [PubMed]
  353. Kumar, A.; Ehrenshaft, M.; Tokar, E.J.; Mason, R.P.; Sinha, B.K. Nitric oxide inhibits topoisomerase II activity and induces resistance to topoisomerase II-poisons in human tumor cells. Biochim. Biophys. Acta 2016, 1860, 1519–1527. [Google Scholar] [CrossRef] [PubMed]
  354. Zhang, A.; Lyu, Y.L.; Lin, C.-P.; Zhou, N.; Azarova, A.M.; Wood, L.M.; Liu, L.F. A protease pathway for the repair of topoisomerase II–DNA covalent complexes. J. Biol. Chem. 2006, 281, 35997–36003. [Google Scholar] [CrossRef] [PubMed]
  355. Ledesma, F.C.; El Khamisy, S.F.; Zuma, M.C.; Osborn, K.; Caldecott, K.W. A human 5′-tyrosyl DNA phosphodiesterase that repairs topoisomerase-mediated DNA damage. Nature 2009, 461, 674–678. [Google Scholar] [CrossRef] [PubMed]
  356. Aparicio, T.; Baer, R.; Gottesman, M.; Gautier, J. MRN, CtIP, and BRCA1 mediate repair of topoisomerase II–DNA adducts. J. Cell Biol. 2016, 212, 399–408. [Google Scholar] [CrossRef] [PubMed]
  357. Quennet, V.; Beucher, A.; Barton, O.; Takeda, S.; Löbrich, M. CtIP and MRN promote non-homologous end-joining of etoposide-induced DNA double-strand breaks in G1. Nucleic Acids Res. 2011, 39, 2144–2152. [Google Scholar] [CrossRef] [PubMed]
  358. Adachi, N.; Suzuki, H.; Iiizumi, S.; Koyama, H. Hypersensitivity of nonhomologous DNA end-joining mutants to VP-16 and ICRF-193: Implications for the repair of topoisomerase II-mediated DNA damage. J. Biol. Chem. 2003, 278, 35897–35902. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic view of the most common lesions causing replication stress. In the scheme, several important replication stress (RS) inducing factors are illustrated: intra-strand crosslink (ISC), inter-strand crosslink (ICL), alkylated/modified base (Me) and inhibition of replication related enzymes. Compounds further described in the review are marked by red colour. RNR: ribonucleotide reductase; DNA pol.: DNA polymerase; TopoI: topoisomerase I; TopoII: topoisomerase II; APH: aphidicolin; HU: hydroxyurea; CPT: camptothecin; ETP: etoposide; cisPt: cisplatin; dATP: deoxyadenosine triphosphate; dTTP: deoxythymidine triphosphate; dCTP: deoxycytidine triphospahte; dGTP: deoxyguanine triphosphate.
Figure 1. Schematic view of the most common lesions causing replication stress. In the scheme, several important replication stress (RS) inducing factors are illustrated: intra-strand crosslink (ISC), inter-strand crosslink (ICL), alkylated/modified base (Me) and inhibition of replication related enzymes. Compounds further described in the review are marked by red colour. RNR: ribonucleotide reductase; DNA pol.: DNA polymerase; TopoI: topoisomerase I; TopoII: topoisomerase II; APH: aphidicolin; HU: hydroxyurea; CPT: camptothecin; ETP: etoposide; cisPt: cisplatin; dATP: deoxyadenosine triphosphate; dTTP: deoxythymidine triphosphate; dCTP: deoxycytidine triphospahte; dGTP: deoxyguanine triphosphate.
Biomolecules 07 00019 g001
Figure 2. Cisplatin structure.
Figure 2. Cisplatin structure.
Biomolecules 07 00019 g002
Figure 3. Aphidicolin structure.
Figure 3. Aphidicolin structure.
Biomolecules 07 00019 g003
Figure 4. Hydroxyurea structure.
Figure 4. Hydroxyurea structure.
Biomolecules 07 00019 g004
Figure 5. Camptothecin structure.
Figure 5. Camptothecin structure.
Biomolecules 07 00019 g005
Figure 6. Etoposide structure.
Figure 6. Etoposide structure.
Biomolecules 07 00019 g006
Table 1. Effects of various cisplatin treatments in vitro.
Table 1. Effects of various cisplatin treatments in vitro.
ConcentrationIncubation TimeObserved EffectCell LineReference
300 μM2 hincrease in polyADP ribosylationO-342 rat ovarian tumour cells[137]
100 μM2 h before IRsensitization to γ-radiationhypoxic V-79
Chinese hamster cells
[138]
100 μM2 hincrease in polyADP ribosylationCV-l monkey cells[139]
<20 μg/mL
(<66 μM)
5 hblock of rRNA synthesis
block of DNA replication
Hela[140]
15 μM1 hinduction of SCE (sister chromatid exchange) decreased cell survival6 primary human tumour cell culture[141]
10–30 μM24 h, 48 hinduction of apoptosis224 (melanoma cells)
HCT116
[142]
10 μM24 hincrease in antiapoptotic Bcl-2 mRNA synthesis (regulated by PKC and Akt2)KLE
HEC-1-A
Ishikawa
MCF-7
[143]
2–10 μM72 hinduction of apoptosis224 (melanoma cells)
HCT116
[142]
5 μM24 hincrease in p53 stability
activation of ATR increased p53(ser15) phosphorylation
A2780[144]
5 μM24 hactivation of p21
activation of CHK2 increased p53(ser20) phosphorylation
HCT116[144]
5 μM24 hinduction of mitochondrial reactive oxygen species (ROS) responseA549
PC3
MEF
[143]
2 μM24 hG2/M arrest, subapoptic damageMSC[145]
>2 μM24 hdecreased proliferation rate
induction of apoptosis
TGCT H12.1
TGCT 2102EP
[145]
1–4 μg/mL2 hblock of DNA synthesisL1210/0 cells[146]
block of transcription
G2 arrest
apoptosis
2 μg/mL48 h
144 h, 168 h
inhibition of mtDNA replication
inhibition of mitochondrial genes transcription
Dorsal root
ganglion (DRG) sensory neurons
[147]
1 μg/mL2 htransient G2 arrestHela[148]
3.0 μM4 h beforeblock of NHEJA2780[138]
0.2–0.8 μMIR 0.5 GycisPt-IR synergistic interactionMO59J
MO59K
[138]
4 h
1–2.5 μM24 h–48 hblock of DNA replication followed by cell apoptosisHela[149]
0.3–1 μMovernightinhibition of RNA polymerase II-dependent transcriptionHela
XPF
[144]
0.6 μM2 h90% reduction in clonogenic capacity detected after 7 days
CHK1 phosphorylation causing CHK1 dependent S phase arrest
Hela[148]
0.5 μM24 h
48 h
loss of telomeres (TEL), or TEL repeats
cell death
Hela[139]
ATR: Ataxia telangiectasia Rad3-related; Bcl: B-cell lymphoma; CHK1: checkpoint kinase 1; CHK2: checkpoint kinase 2; IR: ionizing radiation; mtDNA: mitochondrial DNA; NHEJ: non-homologous end-joining; PKC: protein kinase C; polyADP: poly adenosine diphosphate; rRNA: ribosomal RNA.
Table 2. Effects of various aphidicolin treatments in vitro.
Table 2. Effects of various aphidicolin treatments in vitro.
ConcentrationIncubation TimeObserved EffectCell LineReference
0.2 mM16 h, 10 hformation of anaphase bridges and micronucleiHeLa[204]
30 μM6 hstalled replication forksHCT116[205]
30 μM6 hstalled replication forksPD20 cells
Bloom syndrome cells
[206]
5 μg/mL
(14.3 μM)
4 hDNA repair synthesis inhibition
sensitization towards TNF treatment
L929
ovarian cancer cells A2780
[202]
5 μg/mL
(14.3 μM)
2–8 hS phase arrest
kinetics and mechanism study
RKO
293T
MEF
[207]
2.5 μg/mL
(7.15 μM)
1 hinhibition of DNA synthesis and DNA repairNormal and XPA deficient human fibroblasts[203]
10 μM15 hcell cycle synchronisation at the G1/S boundaryREF-52
HeLa
[208]
5–25 μM24 hinhibition of replicative polymerasesWerner syndrome cells
Bloom syndrome cells
[209,210]
1 μM1–24 hCFS inductionHEK293T[210]
1 μM24 hCFS inductionMEF
HeLa
[211]
0.5 μM2 htransient attenuation of DNA synthesis,DT40[212]
0.1 μM24 hstudy of chromosome integrity and replication
0.4 μM24 hCFS inductionU-2 OS[213]
0.1 μM
0.2 μM
16 hreplication stress observed on telomereshESC (UCSF4)[214]
0.2 μM2 weeksirreversible senescence inductionREF-52[74]
0.2 μM24 hCFS inductionBJ-hTERT[215]
0.05 μM
0.4 μM
24 hCFS inductionWerner syndrome fibroblasts
AG11395 cells
[216]
0.3 μM48 hincreased incidence of mitotic extra chromosomes
replication stress
V79 hamster cell lines[217]
0.3 μM72 hreplication stressHuman fibroblasts HGMDFN090[199]
2 μg/mLnot indicatedreplication blockBJ
BJ-tert
HMEC
[197]
0.2 μM7–24 hcell synchronizationHeLa[184]
CFS: common fragile site; TNF: tumour necrosis factor.
Table 3. Effects of various hydroxyurea treatments in vitro.
Table 3. Effects of various hydroxyurea treatments in vitro.
ConcentrationIncubation TimeEffectCell LineReference
200 mM2 hreplication blockyeast cells[240]
10–200 mM3 hreplication block
replication fork (RF) restart
yeast cells[241]
5 mM1 hreplication blockHEK293[242]
2 mM3 hreplication block
50 μM–5 mM40 min–2 hreplication stress293T
mouse ES cells
[243]
2 mM1 h, 24 hreplication stress
replication block
HCC1937
MCF7
[244]
2 mM16 hreplication blockHEK293[245]
2 mM24 hDNA damage induction during S phaseU-2 OS
293T
[246]
2 mM15 hreplication block
cell cycle synchronisation at the G1/S boundary
REF-52
HeLa
[208]
2 mM5 hdNTP depletionREF52[74]
2 mM3 hchromosomal aberrations
FANCD2 pathway involvement
lymphoblastoid cell lines[247]
1 mMovernightreplication blockMCF7[248]
0.5 mM5 h–10 hreplication blockU-2OS[249]
2 mM2 h–24 hreplication block
0.5 mM90 minnucleotides depletion
stalled RF w/o DSBs formation
MEF[250]
0.1–0.5 mM2 h–72 hγ-globin gene expressionK562[251]
0.1–0.5 mM2 h–8 hreplication stressPC3[252]
0.2–0.4 mM4 dayscell differentiation
ERK signalling pathway inhibition
p38 signal transduction activation
K562[253]
0.3 mM10 daysmicrosatellite instability upon FANCJ depletionGM08402
HeLa
PD20F
[254]
0.15–0.2 mM2 weeksirreversible senescence inductionREF-52[74]
0.2 mM2 h–7 hreplication stressMEF[255]
0.15 mM2 hp53 activationREF52[74]
50–200 μM20 hHIF1 induction
eNOS induction
HUVEC[256]
25–200 μM72 hinduction of apoptosisAML cell lines (MV4-11, OCI-AML3, MOLM-13, and HL-60)[257]
5 μM–0.5 mM48 hreplication stressV79 hamster cells[217]
2 μM12 hreplication stressH1299[258]
dNTP: deoxynucleotide triphosphate; DSBs: double-strand breaks; eNOS: endothelial nitric oxide synthase; ERK: extracellular signal-regulated kinases; FANCD2: Fanconi anaemia complementation group D2; FANCJ: Fanconi anaemia complementation group J; HIF: hypoxia induced factor 1.
Table 4. Effects of various camptothecin treatments in vitro.
Table 4. Effects of various camptothecin treatments in vitro.
ConcentrationIncubation TimeObserved EffectCell LineReference
20 μM30 minDNA fragmentation in G1 and S phase cellsHela[292]
10 μM24 hincrease in cell sensitivity to TRAIL-mediated apoptosisHep3B[293]
10 μM4 hformation of replication mediated DNA DSBsHT29[294]
5 μM60 mininhibition of RNA synthesisCSA[295]
1 μM60 mininhibition of DNA synthesisCSB[296]
1 μM60 minreplication block
DSB formation
cell death
U2OS[297]
1 μM60 minformation of stabilised TopoI-cc complex
DSB formation phosphorylation of CHK1 (S317) CHK2 (T68), RPA (S4/S8)
HCT116[294]
1 μM60 mininhibition of DNA replication
suggested DNA DSB formation
L1210 mouse lymphoblastic leukaemic cells[293]
200 nM–1 μM50 minDSB formationCSB[298]
100 nM–10 nM60 minDSB formationHCT116[299]
25 nM60 mincheckpoint activation (ATM-CHK2, ATR-CHK1)
replication fork stalling
replication fork reversal formation of specific DNA structures
U-2O-S[300]
10 nM–100 nM60 mininhibition of EIAV (equine infectious anemia virus) replicationCF2Th[295]
10 nM–20 nM60 mininhibition of HIV-1 replication
block of viral protein expression
H9[281]
6 nM6 haccumulation of cells in early S phaseNormal lymphocytes[296]
24 hapoptosis, DNA fragmentationMOLT-4
6.25 nM48 hspecific suppression of oral cancer cells growthKB oral cancer cells[281]
2.5 nM48 hincrease in SCE upon depletion of Fbh1 helicaseBJ[281]
ATM: Ataxia telangiectasia mutated; HIV: Human immunodeficiency virus; RPA: replication protein A; SCE: sister chromatid exchange; TopoI-cc: Topoisomerase I cleavage complex; TRAIL: TNF alpha related apoptosis inducing ligand, TNF: tumour necrosis factor.
Table 5. Effects of various etoposide treatments in vitro.
Table 5. Effects of various etoposide treatments in vitro.
ConcentrationIncubation TimeEffectCell LineReferences
up to 450 µM40 minSSB and DSB formation, induction of H2AX phosphorylation with slow kineticsSV-40 transformed human fibroblasts
G361
[323]
1–100 µM30 minformation of TopoII-blocked DSBs, activation of ATM-mediated repairMEF
HEK293T
BJ1
AT
[324]
2–100 μM6 h–48 hsenescence, apoptosis
induction of p53 response
HepG2
U2OS
[325]
2–100 μM1–3 hdisassembly of replication factoriesAT1 BR
AT3 BR
HeLa
[315]
50–100 μM3–6 h/16 hapoptosis (activation of intrinsic (mitochondrial) pathway)Hela
HCT116
[326]
50 μM15 hapoptosisBJAB
Hut78
[327]
50 μM48 hgrowth arrest (accumulation of cells at G2/M boundary)
induction of p53 response
MCF-7
ZR75-1
T-47D
[328]
25 µM1 hSSB and DSB formation
γH2AX, pATM, pDNA-PKcs, MDC1 foci formation
persisting DSBs
cell death
HeLa
HCT116
[329]
20 μM16 hincrease in γH2AX levels
reduction of proliferation rate (accumulation of cells in S and G2/M boundary)
U2OS[330]
20 μM1 hrepairable DSBsHEK293T
COS-7
BJ-hTERT
H1299
[331,332]
16 hirrepairable DSBs, ATM-dependent HIC1 SUMOylation, induction of p53-dependent apoptotic response
20 µM1–5 hapoptosisA549
HeLa,
T24
[333]
10 μM1 hDNA damage inductionA549[334]
1–10 μM48 hapoptosisHCC1937
BT-549
[335]
8 μM1 hinduction of p53 response,SH-SY-5Y
SH-EP1
[336]
0.75–3 µM72 hsenescence, apoptosisA549[337]
0.75 μM24 hcell cycle arrest in G2/M phase, DNA damage induction,
induction of p53 response
MSC
TGCT H12.1
TGCT 2102EP
[145]
DSBs: Double strand breaks; HIC1: Hypermethylated In Cancer 1; MDC1: Mediator of DNA Checkpoint 1; pATM: phosphorylated Ataxia elangiectasia Mutated; pDNA-PKcs: phosphorylated DNA Protein Kinase catalytic subunit; SSB: single-strand DNA break; TopoII: Topoisomerase II.

Share and Cite

MDPI and ACS Style

Vesela, E.; Chroma, K.; Turi, Z.; Mistrik, M. Common Chemical Inductors of Replication Stress: Focus on Cell‐Based Studies. Biomolecules 2017, 7, 19. https://doi.org/10.3390/biom7010019

AMA Style

Vesela E, Chroma K, Turi Z, Mistrik M. Common Chemical Inductors of Replication Stress: Focus on Cell‐Based Studies. Biomolecules. 2017; 7(1):19. https://doi.org/10.3390/biom7010019

Chicago/Turabian Style

Vesela, Eva, Katarina Chroma, Zsofia Turi, and Martin Mistrik. 2017. "Common Chemical Inductors of Replication Stress: Focus on Cell‐Based Studies" Biomolecules 7, no. 1: 19. https://doi.org/10.3390/biom7010019

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop