Next Article in Journal
A Novel Model for Simultaneous Evaluation of Hyperoxia-Mediated Brain and Lung Injury in Neonatal Rats
Previous Article in Journal
Clinical Value of Tocilizumab in Reducing Mortality in Refractory Septic Shock in Children with Hematologic and Non-Hematologic Diseases
Previous Article in Special Issue
Old Passengers as New Drivers: Chromosomal Passenger Proteins Engage in Translesion Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Control of Replication Stress Response by Cytosolic Fe-S Cluster Assembly (CIA) Machinery

Dipartimento di Biotecnologie e Bioscienze, Università degli Studi di Milano-Bicocca, 20126 Milano, Italy
*
Author to whom correspondence should be addressed.
These authors equally contribute to this work.
Cells 2025, 14(6), 442; https://doi.org/10.3390/cells14060442
Submission received: 19 February 2025 / Revised: 11 March 2025 / Accepted: 14 March 2025 / Published: 16 March 2025

Abstract

:
Accurate DNA replication is essential for the maintenance of genome stability and the generation of healthy offspring. When DNA replication is challenged, signals accumulate at blocked replication forks that elicit a multifaceted cellular response, orchestrating DNA replication, DNA repair and cell cycle progression. This replication stress response promotes the recovery of DNA replication, maintaining chromosome integrity and preventing mutations. Defects in this response are linked to heightened genetic instability, which contributes to tumorigenesis and genetic disorders. Iron–sulfur (Fe-S) clusters are emerging as important cofactors in supporting the response to replication stress. These clusters are assembled and delivered to target proteins that function in the cytosol and nucleus via the conserved cytosolic Fe-S cluster assembly (CIA) machinery and the CIA targeting complex. This review summarizes recent advances in understanding the structure and function of the CIA machinery in yeast and mammals, emphasizing the critical role of Fe-S clusters in the replication stress response.

Graphical Abstract

1. Introduction

DNA replication is crucial for all living organisms to transmit genetic information to their progeny. During DNA replication, genome integrity is challenged by both endogenous and exogenous factors, including polymerase-blocking DNA lesions, DNA secondary structures, DNA-binding proteins and DNA-RNA hybrids, all of which can compromise the faithful transmission of genetic material to offspring [1,2]. In addition, alterations in origin firing activation, the depletion of replication factors or low deoxyribonucleotide triphosphate (dNTP) levels can induce replication stress, which hinders replication fork progression [3,4]. Both DNA damage and replication stress can lead to genome instability, which is a hallmark of neurodegenerative diseases and pre-neoplastic lesions, as well as a driver of carcinogenesis [5,6].
Iron–sulfur (Fe–S) clusters are increasingly recognized as essential cofactors in the cellular response to replication stress. They have been identified in various enzymes crucial for DNA metabolism in both yeast and humans, including DNA polymerases, DNA helicases and DNA repair enzymes [7]. While the precise functions of these clusters are not yet fully understood for all Fe–S-containing proteins, they are known to frequently act as enzymatic cofactors and play a role in sensing and responding to the redox conditions within the cell. In some cases, they can also contribute to the structural stability of protein complexes [7,8].
Baker’s yeast has been widely used as a powerful model system to dissect the molecular mechanisms of the replication stress response and to study Fe–S cluster biogenesis in living cells, including Fe-S cluster transfer to target proteins and their impact on protein activity. Here, we review recent advancements in understanding the role of Fe–S clusters in modulating the cellular response to replication stress. Our focus is primarily on findings obtained in Saccharomyces cerevisiae, but we also highlight both similarities and differences between yeast and mammals. Mutations in genes encoding proteins involved in the replication stress response, as well as in factors required for Fe–S cluster biogenesis, have been identified in various cancers and genetic disorders. Therefore, investigating these pathways and their interplays could provide valuable insights into disease pathogenesis and cancer development, potentially guiding new therapeutic strategies.

2. The Cellular Response to Replication Stress

2.1. DNA Replication and DNA Polymerases

DNA replication requires several proteins, including DNA helicases and DNA polymerases, which are assembled into a structure called the replisome, containing all the activities necessary to copy DNA strands with high fidelity [9]. The replisome contains the Cdc45/Mcm2-7-GINS (CMG) replicative helicase that unwinds the parental chromosomes, generating DNA replication forks. At the replication forks, the DNA polymerases ε (Pol ε) and δ (Pol δ) catalyze DNA synthesis on the leading and lagging strands, respectively. Nevertheless, these DNA polymerases are not able to start de novo synthesis, for which they need the action of the Pol α-DNA primase complex that synthesizes RNA-DNA primers with a free 3’ hydroxyl group. The Pol α-primase complex is connected to the replisome, and specifically to the GINS complex, via Ctf4 in both yeast and humans (Figure 1) [10,11,12,13].
Pol δ is composed of the catalytic subunit Pol3/p125 and the accessory subunits Pol31 and Pol32 in yeast, and p50, p66 and p12 in humans [8]. Yeast Pol3 has a proofreading exonuclease activity, which belongs also to mammalian p125, and mutations in POL3 increase the mutation rate [14]. In addition, Pol3 binds a Fe-S cluster, which promotes its association with Pol31 and Pol32 and the formation of the active complex [15]. Pol ε is composed in yeast of the catalytic subunit Pol2 and the accessory subunits Dpb2, Dpb3 and Dpb4 [16,17]. The C-terminal domain (CTD) of Pol2 binds Dpb2, which targets the polymerase toward the replisome [18], whereas Dpb3 and Dpb4 interact with the central portion of Pol2 and participate in DNA binding and chromatin remodeling [19]. In human cells, Pol ε is composed of the catalytic subunit p261 and the accessory subunits p59, p17 and p12 [8]. Pol α is present in eukaryotic cells in tight association with the DNA primase complex. Pol α is composed of the catalytic subunit Pol1/p180 and the accessory subunit Pol12/p70, while the primase complex is composed of Pri1/p49, with catalytic activity, and the accessory subunit Pri2/p58, which contains a Fe-S cluster in the CTD [20,21]. The primase complex synthesizes the RNA primer, whereas Pol1/p180 elongates the RNA primer by adding dNTPs, thus creating a suitable substrate for the action of Pol ε and Pol δ.

2.2. Replication Stress and the Replication Checkpoint

Conditions known as replication stress, which include the presence of DNA-binding proteins, depletion of replication factors or low dNTP levels, hamper the progression of DNA replication (Figure 1) [3,4]. Dysfunctional replication forks are unable to complete DNA synthesis, and cells that are unable to complete DNA replication can accumulate DNA lesions and/or chromosome rearrangements [22]. Cells cope with replication stress by activating the S phase checkpoint, a complex cellular response that couples a delay in S phase progression with DNA repair and the recovery of DNA replication [23]. This response includes the regulation of replication origin firing and of fork architecture, as well as a reorganization of the replication machinery that depends on the source of stress [4,24]. On one hand, small DNA lesions that pass through the replisome and block DNA replication could result in a physical uncoupling between replicative helicases and DNA polymerases, generating a single-stranded DNA (ssDNA) region that is coated by Replication Protein A (RPA) [25]. On the other hand, bulky impediments, such as protein–DNA complexes, interstrand cross-links or DNA adducts, hinder the progression of replication forks by blocking the unwinding of parental DNA strands, generating a stalled replication fork [1,2]. In addition, most cytotoxic lesions such as DNA double-strand breaks (DSBs) can occur when replication forks encounter DNA lesions or ssDNA gaps, generating one-ended DSBs [26].
The presence of either long ssDNA stretches or DSBs activates the checkpoint. The checkpoint is initiated by the protein kinases Mec1/ATR and Tel1/ATM, which recognize long ssDNA stretches and DSBs, respectively [27]. In turn, Mec1/ATR and Tel1/ATM block cell cycle progression through the phosphorylation of the downstream checkpoint kinases Rad53/CHK2 and Chk1/CHK1, whose activation requires the action of the mediators Mrc1/CLASPIN or Rad9/53BP1 [27]. During S phase, Mec1/ATR recognizes ssDNA covered by RPA, which is a common intermediate in many physiological pathways such as DNA replication and transcription. Obstacles to replication fork progression cause fork stalling and the generation of long ssDNA stretches due to nucleolytic processing and the uncoupling between the replicative helicase and polymerase (Figure 1) [2]. Mec1/ATR activation during S phase requires higher amounts of RPA-coated ssDNA compared to what is needed to trigger checkpoint activation in different cell cycle phases, such as G1 or G2. This suggests the presence of a threshold for checkpoint activation during S phase, ensuring that the ssDNA typically produced at functional replication forks is not sufficient to induce a DNA damage response [28]. The recognition of RPA-coated ssDNA by Mec1/ATR depends on the Mec1/ATR-interacting protein, called Ddc2/ATRIP, but the full activation of the kinase also requires the 9-1-1 complex (known also as PCNA-like), which is formed by Ddc1, Rad17 and Mec3 in yeast and RAD9, RAD1 and HUS1 in humans. This complex is loaded on the DNA by Rad24-Rfc2-5 and activates Mec1/ATR, recruiting Dpb11/TopBP1 [29]. The activation of Rad53/CHK2 by Mec1/ATR in response to replication stress is not mediated by the DNA damage checkpoint protein Rad9/53BP1 but requires the mediator Mrc1/CLASPIN. Mrc1/CLASPIN is a component of the replisome; it interacts with both DNA Pol ε and Mcm6 and forms a complex with Tof1/TIMELESS and Csm3/TIPIN, which contributes to the stabilization of the replisome on DNA [30]. Mec1/ATR phosphorylates Mrc1/CLASPIN, which in turn activates Rad53/CHK2. Chk1 in yeast has a relatively minor role in the S phase checkpoint and differs from mammalian CHK1, which is known as the functional counterpart of yeast Rad53 [29]. Upon activation, the effector kinases Rad53 and CHK1 regulate several processes in order to fix DNA damage. These processes include the inhibition of late origin firing, the arrest of the cell cycle, the regulation of dNTP production, the promotion of DNA damage response gene transcription and the remodeling of the replication forks (Figure 1) [27,29]. The suppression of late origin firing and the inhibition of cell cycle progression supply extra time for the resolution of stress and allow the cell to complete DNA replication in the proximity of blocked replication forks. In humans, CHK1 negatively regulates CDK-dependent phosphorylation events at replication origins, blocking the loading of pre-initiation complex factors [31], while in yeast, Rad53 phosphorylates and inactivates the initiation factors Sld3 and Dbf4, thus blocking origin firing [32]. Checkpoint activation also prevents the entry of cells with under-replicated or damaged DNA into mitosis. In fact, CHK1 phosphorylates and inactivates CDC25, preventing G2-M transition. Metaphase–anaphase transition in yeast is instead inhibited by stabilizing the securin Pds1 [29].

2.3. The Control of dNTPs by the Replication Checkpoint

The homeostasis of the dNTP pool depends on the coordinated functions of three main pathways—de novo biosynthesis, nucleoside salvage and nucleotide catabolism—and is tightly regulated by complex feedback networks [33,34,35]. The enzyme ribonucleotide reductase (RNR) catalyzes the rate-limiting step in de novo biosynthesis pathways, which are the main pathways involved in dNTP production in S. cerevisiae, although a pyrimidine deoxyribonucleotide salvage pathway has also been described. This makes yeast an excellent model system for studying de novo dNTP synthesis as well as RNR function and regulation.
RNR catalyzes the reduction of ribonucleoside diphosphates (rNDPs) to deoxyribonucleoside diphosphates (dNDPs), the first step in dNTP biosynthesis. Its specific activity on different rNDPs is subject to sophisticated allosteric effector controls that are crucial to maintain a balanced dNTP pool. This control is ensured by two allosteric control sites located into the R1 subunit, which also contains the catalytic site [33,34]. In eukaryotes, RNR is a class I heterotetrameric enzyme, which comprises a large R1 subunit, composed of a homodimer of α peptides (α2), named Rnr1 in S. cerevisiae and RRM1 in mammals, and a small R2 subunit, which is composed of two identical β peptides (RRM2) in mammals (β2) and of two different proteins denoted Rnr2 and Rnr4 in yeast (β-β’) [33,34]. Yeast cells also express a second RNR large subunit, called Rnr3. Rnr3 is expressed at very low levels compared to the other three RNR subunits and is not essential for cell viability [36,37]. Despite Rnr3 exhibiting low endogenous activity alone, its activity increases in association with Rnr1, and RNR3 overexpression rescues the lethality of rnr1 null mutants [36,37]. The low levels of Rnr3 are due to the repression of RNR3 transcription by the concerted action of the chromatin remodeler ISW2 and the Ssn6-Tup1 repression complex, which are recruited to RNR2, RNR3 and RNR4 promoters by the DNA-binding protein Crt1 [38]. An additional RNR repression mechanism is mediated by the protein Sml1, which binds the yeast R1 subunit and inhibits RNR activity when DNA synthesis is not required [39]. Moreover, an alternative R2 subunit (RRM2B or p53R2) is expressed under the control of p53 in humans and has important functions in regulating mitochondrial DNA replication [34]. RRM2B/p53R2 transcription is also induced in response to DNA damage. RRM2B/p53R2 forms an active RNR holoenzyme in association with RRM1, which supports DNA repair [29]. Once DNA replication and/or repair is completed, RRM2/p53R2 undergoes degradation through ubiquitin-dependent mechanisms involving two E3 ubiquitin ligase complexes: the Skp1/Cullin/F-box (SCF) complex and the anaphase-promoting complex (APC) [40].
In response to DNA damage and replication stress, RNR expression and activity are induced to upregulate the dNTP pool. Yeast Mec1 activates Rad53, which in turn activates the Dun1 effector kinase by phosphorylation. Then, Dun1 enhances RNR function by phosphorylating and inhibiting Crt1 and Sml1 [38,41]. Mec1 and Rad53 regulate the dNTP pool even in the absence of DNA damage, and this function is essential for cell viability. In fact, cells lacking Mec1 or Rad53 are unviable, but this lethality can be rescued by increasing the dNTP pool through the deletion of SML1 or the overexpression of RNR genes [39].

2.4. Fork Remodeling After Replication Stress

Eukaryotic cells have developed sophisticated mechanisms to rescue nonfunctional replication forks and promote fork recovery. These mechanisms depend on the type of obstacle and upon the organization of the arrested forks. Arrested forks can restart by merging forks emanating from dormant origins. This type of origin is abundant in the cell, and it is activated only in the presence of replication stress [42]. Replication forks that are unable to restart from dormant origins could collapse for the release of the replisome from DNA, thus generating structures recognized as DNA damage. Moreover, replication forks coming across DNA lesions can skip them using several strategies with minimal effect on DNA synthesis [43]. Replisomes could omit the lesion through repriming mechanisms that restart DNA synthesis downstream of the lesion [44]. Indeed, replication fork progression could be allowed by DNA polymerases, known as translesion DNA polymerases, which copy the genetic information through a lesioned DNA strand, with lower accuracy compared to the canonical DNA polymerases [45]. Alternatively, the rescue of the replication fork after repriming could involve recombination mechanisms, such as template switching, which requires the action of recombination factors to restore normal DNA replication [24]. An additional key mechanism for stabilizing and restarting replication forks includes the paring of newly synthesized DNA strands and the creation of a four-way structure like a Holliday junction (HJ) [46]. This process, known as fork reversal, was first recognized in yeast as a detrimental conversion that occurs when the Mec1/ATR pathway is abrogated [47]. However, reversed forks have been detected in different species in response to genotoxic agents and have been proposed to stabilize replication forks during the replication stress response. Fork reversal and the subsequent recovery of the reversed fork involve the action of several factors such as nucleases, helicases and recombination factors [24]. For example, the recombinase RAD51 facilitates fork reversal by promoting the coordinated paring of the two newly synthesized DNA strands [48], while the human RECQ1 helicase is essential for the subsequent fork restart [24]. Moreover, different DNA translocases, such as Rad5 in yeast, may also play a role in promoting fork reversal, although their precise function must be deeply investigated [49]. Moreover, RAD51 is involved in the restart and repair of blocked or broken replication forks through a process called HR-mediated fork restart. This mechanism is applied to forks that encounter replication-associated one-ended DSBs or uncoupled or reversed forks after the cleavage of one strand by structure-specific endonucleases such as Mus81 [50].

3. The Iron–Sulfur Machinery

3.1. Structure and Function of Fe-S Clusters

Fe-S clusters are present in proteins that regulate a wide array of cellular functions, including DNA metabolism, mitochondrial respiration, lipid metabolism and protein translation [51]. Fe-S clusters are ancient, small, inorganic protein cofactors composed of ferrous (Fe2+) or ferric (Fe3+) iron and sulfur inorganic ions (S2-). Some Fe-S proteins can also contain additional heavy metals such as molybdenum, vanadium or nickel [52]. Fe-S clusters can be found in three different forms: [2Fe-2S]+, [4Fe-4S]2+ and [3Fe-4S]+. In the simplest form [2Fe-2S]+, sulfur and iron ions coordinate four cysteine side chains, thus creating a rhombic structure. [4Fe-4S]2+ is considered a duplication of the [2Fe-2S]+ structure, where iron and sulfur alternately occupy the corners of the cube defined by ion disposition. [3Fe-4S]+ is the rarest form; it has a cuboidal structure in which one corner of the cube is unoccupied. The most common coordination partner for the iron ions is the sulfur atom of cysteine residues. However, histidine, and in rare cases arginine, serine, peptidyl-N and non-protein ligands, may also be used [53,54,55].
No single common consensus motif for binding a Fe-S cluster has been identified. However, some Fe-S proteins exhibit particular sequences, such as the CX4CX2CX ≈ 30C or the CX2CX2CX20–40 motifs. There is a slight preference for proline and tyrosine/phenylalanine residues in the first and second positions after the cysteine, respectively [56]. In addition, a Lys-Tyr-Arg (LYR)-like motif was identified in different Fe-S proteins and was found to mediate the interaction of the HSC20 chaperone with Fe-S client factors involved in the respiratory chain in human cells [57]. The LYR motif is also present in mitochondrial Fe-S proteins of some yeasts, leading to the suggestion that it could drive the insertion of Fe-S clusters in subclasses of Fe-S client proteins [57]. Despite the existence of these Fe-S consensus sequences, many Fe-S proteins do not follow these patterns. This inconsistency makes it challenging to predict and identify Fe-S proteins solely based on their primary sequence information. To address this challenge, several tools have been developed in order to identify novel possible Fe-S proteins [58,59,60].
Fe-S clusters serve various functions, although for some Fe-S proteins, the molecular role of the cofactors remains enigmatic [60]. In certain cases, these clusters can accept electrons from one partner and donate them to other molecules due to the ability of iron ions to switch between reduced and oxidized states. This electron transfer capability is exploited in enzymes involved in nitrogen fixation, photosynthesis and mitochondrial or bacterial respiration [53,61,62]. Fe-S proteins also play roles in non-redox processes: for example, they can function as a Lewis-acid catalyst, they are engaged in radical-mediated catalysis and they can serve as sulfur donors for the synthesis of cofactors [7,61]. In addition, the instability of Fe-S clusters in response to varying iron and oxygen levels makes them suitable for sensing ambient oxygen, gases and iron concentrations, hence regulating gene expression in response to environmental changes [63]. Fe-S clusters have been proposed to be important for the structure and activity of Fe-S proteins involved in DNA metabolism, as well as for their ability to bind DNA and detect DNA damage [41,64]. Finally, in some cases, Fe-S clusters may simply serve a structural role, stabilizing the rest of the polypeptide chain [53].

3.2. Biogenesis of Mitochondrial, Cytosolic and Nuclear Fe-S Cluster Proteins

Fe-S clusters can spontaneously assemble on purified proteins in vitro from ferrous iron and sulfide. However, in living cells, the biogenesis of Fe-S clusters requires complex biochemical assembly systems [7,52]. The biogenesis of Fe-S clusters requires cooperation between the mitochondrial Fe-S cluster (ISC) system and the cytosolic Fe-S protein assembly (CIA) system. The ISC system generates clusters for mitochondrial Fe-S proteins and a sulfur-containing compound that serves in the maturation of cytosolic and nuclear Fe-S proteins. This compound is delivered to the cytosol, where the CIA machinery assembles the Fe-S cluster and inserts it into the target proteins (Figure 2). ISC and CIA factors are conserved in eukaryotes, and most of them are essential for cell viability (Table 1) [7,61].
The ISC pathway starts with the delivery of iron and sulfur ions to the scaffold protein Isu/ISCU. Sulfur is released by the cysteine desulfurase Nfs1 from cysteine residues, while the required electrons are transferred from NADH through the electron transfer chain formed by the ferredoxin Yah1/FDX2 and the ferredoxin reductase Arh1/FDXR [52,53,65,66]. The source of iron is less clear; it is probably imported in the mitochondrion and then delivered to Isu/ISCU thanks to Yfh1/frataxin [53,67,68]. This peptide functions as an iron donor for Fe-S cluster synthesis in vitro, but concrete in vivo evidence of frataxin supplying iron ions is still lacking. Recent data suggest that it also functions as an allosteric regulator, thus increasing the cluster formation rate [54,69]. The ISC pathway also involves chaperones such as Ssq1/HSPA9 and Jac1/HSC20, which stabilize the scaffold protein and facilitate the release of the cluster [70]. The complex formed by ISCA1, ISCA2 and IBA57 assembles a [4Fe-4S]2+ cluster by the reductive coupling of two [2Fe-2S]+ clusters. This complex is also conserved in yeast, where ISCA1 and ISCA2 are called Isa1 and Isa2, respectively. How these three proteins physically and functionally interact with each other is still unclear [54,55,69]. Finally, NFU1/Nfu1 promotes the binding of the cluster to the target mitochondrial apoproteins [54,69,71].
Then, the ISC pathway exports a sulfur-containing compound of still-unknown identity from the mitochondrion to the cytosol for the biogenesis of cytosolic Fe-S clusters (Figure 2). The ABC transporter ABCB7 (Atm1 in yeast) is involved in this export [72]. However, in mammalian cells, the presence in the cytosol of small amounts of several ISC components suggests that maybe Fe-S clusters are not assembled exclusively in the mitochondrion. This aspect is still debated [69].
Cytosolic Fe-S cluster formation begins with the two P-loop NTPases Nbp35/NUBP1 and Cfd1/NUBP2, which form a heterotetrameric protein complex that coordinates different [4Fe-4S]2+ clusters [7,73]. Similarly to the role of ferredoxin in the mitochondrial process, the cytosolic pathway requires an electron transfer chain which involves the diflavin reductase protein Tah18/NDOR1 and the Fe-S protein Dre2/CIAPIN1 (Figure 2) [61,74]. Biogenesis further requires the cytosolic monothiol glutaredoxin Grx3-Grx4 (human PICOT), which promotes Fe-S cluster assembly on Dre2 [75,76]. The second major step of cytosolic Fe-S protein biogenesis involves the release of the newly assembled [4Fe-4S]2+ cluster from Cfd1-Nbp35 and its transfer to the late-acting CIA proteins. This reaction is mediated by Nar1/IOP1, which contains C- and N-terminal motifs with four cysteines each coordinating two [4Fe-4S]2+ clusters. Nar1/IOP1 functions as an adapter at the interface between early (Cfd1-Nbp35, Tah18-Dre2) and late CIA [7]. The late-acting CIA proteins form the CIA targeting complex (Cia1-Cia2-Met18 in S. cerevisiae and CIA1-CIA2B-MMS19 in mammals) and are responsible for the insertion of the cluster into the target proteins (Figure 2). We will now focus on the structure and function of the proteins that form the CIA targeting complex. For further details on the other steps of Fe-S cluster biogenesis, we refer the reader to excellent reviews [54,55,69,77,78].

3.3. The CIA Targeting Complex: Structure and Activity

The CIA targeting complex includes Cia1/CIA1, Cia2/CIA2B and Met18/MMS19 subunits [61]. These components physically interact with a large number of target proteins in the cytoplasm and nucleus. The CIA targeting complex is proposed to mediate apo-target recognition. However, the specific function(s) of the individual subunits within this complex has/have not been clearly defined [79].

3.3.1. Cia1

Cia1 belongs to a large protein family characterized by WD40 repeat domains that form a β-propeller structure with seven pseudo-symmetrically orientated blades around a central axis. This geometry serves as a docking site for Fe-S proteins [80]. Cia1 binds Cia2, while Met18-Cia1 complexes have not been observed. Thus, Cia1’s interaction platform is probably used for target recognition or for the binding of Nar1 [81]. Cia1 arginine 127 in yeast is important for protein–protein interactions because its mutation in glutamic acid (cia1-R127E) affects cell growth and CIA machinery function [80]. The human homologous CIA1 was proposed to also play a role in transcription regulation [82]. Unlike other CIA proteins, yeast Cia1 mainly localizes in the nucleus, suggesting that it may also have a transcriptional role like its human counterpart [53,80,83]. Patients with biallelic loss-of-function mutations in CIA1 develop neuromuscular and multisystemic problems, including muscle weakness, respiratory insufficiency, iron deposition in deep brain nuclei and learning difficulties. These mutated CIA1 variants lead to failure in recruiting Fe-S client proteins, thus resulting in compromising the activity of Fe-S proteins, such as DNA helicases, polymerases and repair enzymes that rely on the CIA complex to acquire their Fe-S cofactors [84]. Cia1/CIA1 appears to be required for Fe-S transfer to all known Fe-S target proteins, while it is dispensable for the maturation of the Fe-S cluster-containing CIA proteins, such as Cfd1, Nar1 and Nbp35 in yeast, and CIAPIN1 in mammals [61,85]. Recent experiments have demonstrated that [2Fe-2S]+ cluster assembly on the target proteins Aft1-Aft2 and Yap5 in yeast, and aldehyde oxidase in mammals, occurs in a CIA-independent manner, thus suggesting that the role of Cia1/CIA1 in target maturation needs further validation [85].

3.3.2. Cia2

Yeast Cia2 is the homolog of human CIA2B (also named FAM96B or MIP18). In addition to CIA2B, human cells express a Cia2-like protein (CIA2A) that is particularly required for the maturation of key regulators of iron homeostasis [86,87]. CIA2B is also important in coordinating ATP and Fe-S cluster homeostasis during mitosis. In fact, it functions as a scaffold protein that recruits Creatin Kinase B and other components of the MMXD complex, which is formed by MMS19, CIA2B and the helicase XPD, and it is involved in chromosome segregation [88,89].
In yeast, Cia2 is the central component of the targeting complex, tethering Met18 to Cia1 [79,81]. It is a small acidic protein that contains five conserved motifs distributed between an intrinsically disordered N-terminal domain and C-terminal domain of unknown function 59 (DUF59). The disordered domain is dispensable for the binding of the other subunits of the targeting complex in vitro, but in vivo, the deletion of the N-terminal domain negatively impacts CIA function. In fact, the deletion of the first 102 amino acids in yeast Cia2 (cia2-Δ102) is sufficient to support cell viability but causes compromised functionality and a diminished ability to support CIA target maturation. This is possibly due to the lower stability of this variant in vivo [79]. The C-terminal region of Cia2 contains the DUF59 domain (Motif 3 and Motif 4) and about 40 amino acids that form Motif 5. Motif 5 is unique in eukaryotic DUF59 proteins, indicating its potential significance in the formation of the targeting complex [61,79]. In Motif 5, glutamate 208 has been demonstrated to be important for Cia1-Cia2 interaction but not for interaction with Met18 in vitro. Mutations of this residue in glycine or alanine do not impair Cia2 function. However, in vivo, they destabilize the Cia1-Cia2 complex, which in turn affects the amount of the Met18-Cia2 complex that can be detected with biochemical assays. They also cause a diminished activity of cytosolic Fe-S cluster proteins [79,90]. Cysteine 161 in yeast and 93 in humans (Motif 4) is critical for the function of Cia2. Mutations of this cysteine cause cell lethality and dominant-negative effects in yeast. This result suggests that Cia2 may engage in protein complexes that are sequestered by the Cia2-C161A mutant variant, thereby disrupting the activity of the wild-type protein and sequestering Met18 and Cia1 in nonfunctional CIA targeting complexes [79,91]. Similarly, mutations in Motif 3 lead to nonfunctional alleles and slow-growing yeast strains. Surprisingly, mutations in Motif 3 and Motif 4 produce no observable perturbations in Cia2′s interaction with the other components of the CIA targeting complex, suggesting that Cia2 uses the DUF59 domain for an additional, yet-unknown function. It has been suggested that Cia2 could play a cluster-carrying role and could interact with the Fe-S cluster as it is inserted into the target apoproteins [79]. Together, these findings highlight how difficult it can be to understand these molecular details and put together in vitro and in vivo results and suggest that Cia2 is not just a mediator of protein–protein interactions, but it likely has an additional, yet-unidentified function in the final cluster insertion step [79].

3.3.3. Met18

Met18/MMS19 is the biggest protein of the CIA targeting complex. It regulates the maturation of several cytosolic Fe-S proteins [7]. In contrast to all the other known CIA proteins, Met18 is not essential for cell viability in yeast, while MMS19 knockout is embryonically lethal in mice [61,92]. A lack of Met18/MMS19 is associated with a multitude of cellular phenotypes, including defects in methionine synthesis (yeast only), impaired chromosome segregation, sensitivity to genotoxic agents or elongated telomeres [61]. Met18 is a 118 kDa protein with an alpha solenoid structure composed entirely of HEAT repeats (Huntington, Elongation Factor 3, Protein phosphatase 2A and TOR kinase), each including two alpha helices that are known to mediate protein–protein interactions and are often part of large protein complexes. In humans, MMS19 lysine residues from helices 45 and 46 are responsible for the interaction between MMS19 and CIA2B, while no direct interaction between MMS19 and CIA1 has been detected. In these helices, the ubiquitination of lysine residues is likely to preclude CIA2B binding due to steric clashes. Besides directing MMS19 to the proteasome for degradation, thus shutting down the CIA pathway, ubiquitination could serve as a regulatory mechanism that immediately inhibits Fe-S protein biogenesis and prevents the further transfer of Fe-S clusters into MMS19-dependent client proteins [93].
The C-terminal region of MMS19 interacts with CIA2B, while the rest of the protein is accessible for interactions with client proteins or early-acting factors of the CIA pathway [93]. Structural studies in yeast have shown that Met18 adopts a different conformation from those observed in the human and mouse homologs [93] that allows it to form hexamers and tetramers in the absence of Cia2. This oligomerization may protect Met18 from degradation when the CIA complex is paused between clients and/or protect it from interactions with non-clients or holo-clients. In these conformations, the residues in the C-terminal region that appear to be involved in binding Cia2 (R1010, R1013, R1020), and those in the N-terminal region that bind the client proteins (for example, R144, K187 and F217) are hidden, indicating that Cia2 cannot bind Met18 without disrupting these higher-order oligomeric states. The addition of Cia2 causes these oligomeric states of Met18 to dissociate and form Met18-Cia2 complexes [94]. The precise molecular mechanisms by which Cia2 achieves this feat are still unclear.

3.4. Assembly of CIA Targeting Complex and Target Maturation

Met18/MMS19, Cia1 and Cia2/CIA2B form the CIA targeting complex, in which Cia2/CIA2B is the central scaffold that links Met18/MMS19 to Cia1 [81,95]. The yeast CIA complex is formed by the docking of two Cia1-Cia22 complexes with Met18, so it comprises one Met18, four Cia2 and two Cia1 polypeptides [81]. In human cells, the CIA complex exists in a dynamic monomer–dimer equilibrium in solution, but further studies are needed to investigate its exact structure [93].
The CIA targeting complex can form various binary and ternary subcomplexes with distinct target specificities. Different human CIA targeting complex components exhibit a high specificity for the maturation of dedicated Fe-S proteins [7]. In fact, proteomic studies show the differential binding of CIA1, CIA2B and MMS19 to the collection of cytosolic and nuclear Fe-S proteins. The ternary CIA targeting complex appears to have the broadest substrate spectrum, mediating the maturation of numerous Fe-S proteins such as dihydropyrimidine dehydrogenase (DPYD), which is involved in nucleotide production, the ATP-binding cassette protein ABCE1 and the helicase XPD. MMS19 plays only a minor role in the maturation of the nucleotide metabolism protein glutamine phosphoribosylpyrophosphate amidotransferase (GPAT), while CIA2B is not crucially required for DNA Pol δ. CIA2A was instead proposed to participate in the regulation of iron homeostasis in human cells, modulating the activity of the iron regulatory proteins IRP1 and IRP2. CIA2A seems to stabilize IRP2 and contribute to the maturation of IRP1, even if no direct interaction has been detected so far between CIA2A and IRP1 [7,86]. The formation of subcomplexes with different specificities appears to be a peculiarity of human cells. However, recent data showed that CIA substrate recognition can be performed by either the full targeting complex or its stable subcomplexes in yeast. For example, Met18-Cia2 can bind the client protein Rad3, whereas the binding of Leu1 requires the full targeting complex [81].
CIA proteins deliver Fe-S clusters to more than thirty thus-far-known client proteins which are structurally distinct. This raises a fundamental question: how do the same three CIA proteins recognize clients of different sizes and shapes while avoiding holo-clients and non-clients [94]? The only similarity detected in different apo-targets is represented by the cysteine residues required to bind the [4Fe-4S]2+ cofactors. However, the CIA complex does not recognize any Fe-S protein in vitro. In fact, it cannot interact with bacterial Fe-S proteins, demonstrating that the cluster-binding cysteine motif is not sufficient for target recognition [81].
Since different CIA subcomplexes interact with distinct subsets of targets, each subunit of the targeting complex contributes to target recognition [81]. Structural analysis of the human CIA complex has revealed the presence of two key interaction sites for recognizing client proteins. One is a high-affinity binding site on CIA1, likely common to all client proteins, while the second is located within the N-terminal domain of MMS19. The latter site facilitates the recruitment of MMS19 client proteins by increasing their binding affinity due to cooperativity between the two sites. The plasticity of MMS19 enables the complex to adapt to diverse molecular shapes of client proteins: for example, MMS19 adopts a different conformation based on whether the CIA complex is bound to DNA2, which is more globular, or to a more elongated protein such as DNA primase. The significant conformational plasticity of MMS19 not only aids in client protein recognition but may also play a role in the mechanism of Fe-S cluster transfer from CIA2B into the targets. Cryo-EM studies with the client protein primase have shown that when DNA primase is bound to the complex, the structure is not in a relevant state for Fe-S cluster transfer, as the primase binding site is distant from the predicted cluster-binding site on CIA2B. This indicates that a conformational change or other factors may be required to facilitate Fe-S cluster transfer from the CIA complex to its client proteins. Complex plasticity may facilitate the transfer of Fe-S clusters, bringing CIA2B closer to the client protein. Alternatively, additional factors might be required to bridge CIA2B and the client proteins during Fe–S cluster transfer [93,94]. Also, in yeast, Met18′s conformational flexibility is key for the recognition of clients of various sizes and shapes because its curvature adapts to its binding partners. Together, these findings suggest that multiple states of the CIA complex exist: a state in which the Fe-S cluster is delivered by Nar1, a state in which the cluster is bound to Cia2 and another in which the cluster is bound to the target protein [94]. Further studies are needed to clarify the structure and activity of these different complexes.

4. Role of CIA Machinery in Response to Replication Stress

The CIA targeting complex is important to safeguard DNA replication under stress conditions. In fact, in both yeast and mammals, Met18/MMS19 defects result in increased sensitivity to the alkylating agent methyl methanesulfonate (MMS) and to UV light, which interfere with DNA replication, but not to X-ray irradiation [96,97,98,99]. Furthermore, yeast cells expressing the Cia2-Δ102 variant are viable but hypersensitive to both MMS and hydroxyurea (HU), which reduces the dNTP pool, suggesting that the Cia2 N-terminal domain is required for the maturation of Fe-S proteins involved in the replication stress response [79].
The hypersensitivity to replication stress of cells with CIA defects could be due to checkpoint defects, as suggested by both the reduced Rad53 phosphorylation observed in HU-treated met18Δ yeast cells and the defective CHK1 phosphorylation detected in HU-treated human HeLa cells [92]. Other studies indicate instead that CIA is not required for checkpoint activation. In fact, the phosphorylation of checkpoint targets Rad53 and CHK1 was observed in both MMS-treated yeast cells lacking Met18 and HU-treated HEK293 human cells with MMS19 depletion, and these phosphorylations were enhanced in CIA-deficient cells compared to the wild type, suggesting that the checkpoint could be hyperactivated in the absence of CIA functions [99,100]. HU treatment causes a prolonged G1 phase and delayed S phase entry in cells with Met18/MMS19 deficiency [92]. Since checkpoint signals in HU are generated during ongoing DNA replication, the reduced checkpoint observed in cells treated in G1 with HU could merely be a consequence of the persistence of these cells in G1. Overall, these findings suggest that the Fe-S machinery is not essential for checkpoint activation but instead facilitates the cellular response to replication stress by minimizing the accumulation of structures that trigger checkpoint activation. In doing so, it enhances the efficiency of the checkpoint pathway in managing replication stress. This hypothesis is supported by the finding that CIA inactivation creates a dependency on the Mec1/ATR pathway for cell survival. In yeast, MET18 deletion reduced the viability of cells carrying a mec1-4 temperature-sensitive allele, although the underlying mechanism was not further investigated [101]. Similarly, the depletion of either MMS19 or CIA2B sensitizes triple-negative breast cancer cells to treatment with inhibitors of ATR or CHK1 proteins [100]. Therefore, Met18/MMS19 supports the functions of the Mec1/ATR pathway in DNA replication control, and this function is shared by other CIA components, suggesting that the CIA complex supports cell viability in the absence of the replication checkpoint by promoting the maturation of Fe-S target proteins.
The Mec1/ATR pathway supports cell viability when DNA replication is threatened by a perturbation of DNA polymerase activity, DNA damage or poor or insufficient availability of dNTPs, but it is also crucial for DNA replication in untreated conditions [102,103]. Recent findings indicate that in yeast, Mec1 is transiently activated every time cells enter S phase in short regions centered on early replication origins, reflecting the activation observed under replication stress [104]. This transient activation of Mec1 in unperturbed S phase is suppressed by increasing the dNTP pool before cells enter S phase, indicating that the scarce availability of dNTPs when DNA replication starts creates a local replication stress that activates Mec1, which in turn stimulates RNR activity and concomitantly prevents replication fork collapse [104]. The CIA complex takes part in the regulation of these processes, likely acting independently of the ATR/Chk1 pathway and supporting the functions of Mec1/ATR both in untreated conditions and under replication stress. Below, we describe how the Fe-S machinery can support the replication stress response by controlling DNA polymerases stability, the recovery of stalled replication forks and the availability of dNTPs. We suggest that the regulation of these aspects by the Fe-S machinery contributes to supporting faithful DNA replication under both normal and stressed conditions (Figure 3).

4.1. CIA and the Control of DNA Polymerases

A clear connection between Fe-S clusters and DNA polymerases in both yeast and humans was made in the first decade of the 2000s, beginning with the identification and characterization of the yeast pol3-13 mutant allele [105]. This allele presents a mutation in the last cysteine within the Pol3 CysB motif and a synthetic lethal phenotype in combination with mutations in both CIA components Met18, Nbp53, Dre2 and Tah18 and checkpoint factors, such as Rad53 [105]. Subsequent works identified Fe-S clusters in both yeast and mammalian DNA primase [15,106]. Since then, the CIA complex has been documented to transfer Fe-S clusters to the Fe-S center of DNA polymerases and primases involved in bulk DNA replication as well as in translesion DNA synthesis, thus regulating their stability and/or their activity [107,108]. However, exploring in vivo the consequences of Fe-S impairment in DNA polymerases remains difficult, as the insertion of the clusters in the replicative DNA Pol α, ε and δ is essential for cell viability.
Replicative polymerases (α, δ and ε), as well as the DNA primase Pri2/p58, contain Fe-S clusters in their CTD. Both in yeast and in humans, mutations of iron-coordinating cysteines in the primase CTD resulted in the disruption of protein folding, indicating that the clusters play a critical structural role [107]. The Pol2 subunit of polymerase ε contains two [4Fe-4S]2+ clusters, one located in the CTD domain, which is stabilized by interaction with the Dpb2 subunit, and the other within the polymerase domain [109,110]. Mutations in the CysX motif, which coordinates the Fe-S cluster in the N-terminal region of Pol2, impair the DNA binding of the complex, thus abrogating both polymerase and exonuclease activities [111].
In addition to increasing protein stability, the Fe-S cluster has been proposed to act as a redox switch that modulates interactions between DNA polymerases and DNA (reviewed in [112]). [4Fe-4S]2+ clusters can be oxidized to the [4Fe-4S]3+ state or reduced to the [4Fe-4S]+ state. Oxidized Fe-S clusters enhance the DNA-binding affinity of their cognate proteins compared to their reduced counterparts. The redox state of the Fe-S center in DNA primase acts as a molecular switch, regulating DNA binding, the interaction of Pol α-primase complex with the RNA primer and the polymerase handoff [113]. The redox state of the Fe-S cluster in the Pol3 CTD instead regulates the rate of DNA synthesis carried out by Pol3 in vitro [114]. Also, the Fe-S cluster of yeast DNA polymerase ε is redox-active, with protein oxidation causing a sharp decrease in DNA synthesis polymerization [115]. The redox switch between oxidized and reduced states can be activated using DNA charge transport chains and can offer a way to slow down or block DNA synthesis in response to oxidative stress [112]. Recent reports demonstrated that besides reducing RNR activity, HU treatment in yeast cells increases the accumulation of reactive oxygen species (ROS), and that in both yeast and human cells ROS cause the disassembly of polymerase complexes into their subunits, thus inhibiting both DNA binding and polymerase activity [116,117]. Importantly, eukaryotic DNA polymerase complexes that contain Fe-S clusters are inhibited by ROS, while prokaryotic DNA polymerases lacking Fe-S clusters are not, suggesting that ROS inhibit polymerase activity by inducing the oxidation of their Fe-S clusters [117]. Therefore, the redox state could modulate the activity of polymerases through the modulation of the Fe-S center.
Mutations affecting Fe-S binding to the Pol3 CTD interfere with Pol δ complex formation and cause cell lethality at high temperatures, suggesting that the Fe-S cluster supports Pol δ structure and function in non-optimal growth conditions [15,105,118]. Furthermore, a mutation in the Pol3 cysteine-rich domain CysB triggers checkpoint activation and increases both DNA damage and genome instability, also in unperturbed growth conditions [119]. The overexpression of RAD51 and RAD52 partially suppresses these defects and increases the viability of pol3-13 mutant cells, indicating that defects in the Pol3 Fe-S center impairs DNA replication and increases DNA damage in S phase, which requires recombination-mediated repair to be fixed [119]. This model is supported by the finding that Pol3 mutations that prevent Fe-S cluster assembly cause a mutator phenotype and the accumulation of complex mutations and small deletions flanked by short homologies, which are compatible with enhanced DNA polymerase slippage and/or template switching events [120]. Importantly, both the rates and the spectra of mutations observed in these pol3 mutants resemble those observed in the absence of Met18, while they are not caused by mutations that decrease Fe-S addition to Pol ε [120]. Consistently, both in yeast and in mammals, Pol δ interacts with Met18/MMS19, as does DNA primase [92]. Furthermore, Met18/MMS19 inactivation severely affects the stability of both Pol α and Pol δ complexes [92]. Therefore, Met18 appears to stimulate the fidelity of DNA synthesis mainly via the regulation of Pol δ processivity.
As the stability of the replisome is important for replication recovery under replication stress conditions and Fe-S clusters stabilize DNA polymerase complexes and their interactions with DNA, CIA can support DNA replication under stress conditions at least in part by stabilizing DNA polymerases at stalled replication forks (Figure 3).

4.2. CIA and the Restart of Replication Forks

Recombination is essential for restoring active replication forks under replication stress and during replication recovery. Additionally, helicases and nucleases play a key role in fork remodeling and the recombination-dependent recovery of stalled forks, contributing significantly to replication fidelity and genome stability [24]. DNA helicases and nucleases are key targets of the CIA targeting complex, and their regulation through Fe-S cluster insertion may facilitate DNA replication recovery under stress conditions. However, the precise roles of the CIA complex and Fe-S clusters in this process remain largely unexplored. The CIA targeting complex inserts Fe-S clusters in the nuclease/helicase DNA2, and the helicases Rad3/XPD, FANCJ, RTEL1 and Chl1/CHLR1/DDX11, which are involved in different mechanisms of DNA repair and in the cohesion of sister chromatids [99,121]. Although both DNA repair and cohesion are essential to maintain genome stability during DNA replication, Dna2 is likely the Fe-S cluster protein mainly involved in fork recovery under replication stress (Figure 3). In yeast, Dna2 contributes to Mec1 activation at stalled replication forks [29]. Furthermore, DNA2 depletion leads to the accumulation of toxic ssDNA structures behind the replication forks that activate the DNA damage response [122]. Finally, Dna2 has been shown to counteract the accumulation of reversed forks and to promote replication fork restart both in yeast and in mammals [123,124]. The contribution of the CIA targeting complex to regulating these Dna2 functions has not been completely defined. However, as the checkpoint is activated in cells lacking Met18 [99,100], Fe-S cluster addition to Dna2 is likely not required for the Dna2-dependent activation of Mec1.
The Dna2 nuclease domain contains four cysteine residues that coordinate a [4Fe-4S]2+ cluster. The coordination of this cluster in DNA2 is unusual, as one of the cysteine residues that binds the cluster is located several hundred amino acids upstream of the other three [125,126]. Dna2 mutations that impair Fe-S assembly in yeast do not affect protein stability or the ability of Dna2 to bind DNA, although they impact the way in which the protein binds DNA, and both Dna2 nuclease and ATPase activities [126]. The loss of the Fe-S cluster in yeast Dna2 also does not impact Dna2 binding to RPA, which recruits Dna2 to 5′-ended ssDNA [126]. The crystal structure of mouse Dna2 indicates that the Fe-S cluster stabilizes a crossover loop within the nuclease domain, forming the base of a cylinder through which ssDNA is channeled [127]. This DNA-binding tunnel undergoes a conformational change following Fe-S cluster addition, which likely facilitates the access of DNA to the nuclease domain. However, Fe-S is not absolutely required for DNA binding, as DNA2 mutant variants carrying mutations in the Fe-S center are still able to bind DNA in vitro. Even in conditions where the protein could bind DNA, the absence of the Fe-S cluster severely impairs the nuclease, helicase and ATPase activities of human DNA2, indicating that the Fe-S cluster specifically affects these Dna2 functions [128]. Fe-S cluster mutants have defects in DNA replication and repair in vivo that correlate in intensity with their effect on catalytic activities in vitro [126]. However, whether and how the Fe-S cluster impacts Dna2 functions at replication forks are still open questions.
Both in yeast and in mammals, Chl1/ChlR1/DDX11 is an ATP-dependent Fe-S cluster DNA helicase required for sister chromatid cohesion. In mammals, the inactivation of the ChlR1/DDX11 gene causes replication stress, renders cells dependent upon DNA damage response genes for survival and causes sensitivity to MMS, mitomycin and camptothecin [129]. CHLR1/DDX11 depletion was also found to sensitize cancer cells to poly ADP-ribose polymerase (PARP) inhibitors and platinum-based chemotherapy and enhances the sensitivity to chemotherapy of cells with BRCA1 or BRCA2 mutations [130]. ChlR1/DDX11 interacts physically with the replication fork component TIMELESS to assist replisome progression [131,132] and promotes ssDNA generation at DNA lesions, thus stimulating both recombination-mediated DNA repair and CHK1 activation [130,133]. These findings suggest that Chl1/ChlR1/DDX11 can take part in replication fork recovery under replication stress (Figure 3). However, how this helicase promotes fork recovery and whether the Fe-S cluster is important for this function are still unknown.
Finally, the Fe-S-mediated regulation of Pol ε could also be involved in replication fork recovery. A switch between polymerase/exonuclease activities of Pol ε was recently documented. This switch was regulated by the checkpoint and was proposed to be important to prevent the extensive processing of stalled forks and promote replication restart [134]. It will be interesting to evaluate whether the Fe-S cluster takes part in this regulation.

4.3. CIA and the Control of dNTPs

The CIA targeting complex can support the cellular response to replication stress by regulating the synthesis of dNTPs (Figure 3). In fact, the depletion of MMS19 results in the deregulation of several metabolites involved in nucleotide production [100]. In both yeast and mammals, CIA inactivation also increases the lethality of cells with replication checkpoint dysfunctions [100,101]. As the replication checkpoint promotes dNTP synthesis [33], the lethality of cells with dysfunctions in both CIA and checkpoint pathways can be due to insufficient or unbalanced dNTP levels [100]. If CIA stimulates dNTP production, CIA inactivation can also exacerbate the reduction in dNTP levels caused by HU treatment. This extreme low availability of dNTPs, which is not sufficient to support DNA replication initiation, can be the cause of the delayed G1/S transition observed in HU-treated met18Δ yeast cells [92].
How can the CIA complex regulate dNTP levels? Iron metabolism impacts dNTP production because RNR is an iron-dependent enzyme. The catalytic process that converts rNDPs to dNTPs requires tight cooperation between both R1 and R2 RNR subunits and a diferric–tyrosyl radical cofactor (Fe3+2-Y∙), which is part of the peptide β in the R2 subunit in both yeast and mammals (Figure 4). This cofactor stabilizes a tyrosyl radical, which is essential for the catalytic process. Substrate binding to R1 initiates an electron transfer chain that involves amino acid residues located in both R1 and R2 subunits and leads to rNDP reduction. After catalysis, the RNR enzyme should be regenerated by the reduction of oxidized cysteines by thioredoxin or glutaredoxin [33,34]. Of the two different β peptides in yeast, only the Rnr2 subunit contains the Fe3+2-Y∙cofactor, while Rnr4 contributes to the assembly of Fe3+2-Y∙ into Rnr2 [135].
The assembly of the Fe3+2-Y β cofactor requires a source of iron and a reducing equivalent. In yeast, the glutaredoxins Grx3 and Grx4, together with the electron donor complex Dre2-Tah18, deliver the iron ion in its reduced state to Rnr2 for the formation of the Fe3+2-Y∙cofactor, supporting RNR activity (Figure 4) [135,136]. Indeed, Grx4 depletion reduces RNR activity due to the inefficient incorporation of iron [136]. Furthermore, Dre2 depletion diminishes both the levels of tyrosyl radical Y∙ and RNR activity, suggesting that Dre2-Tah18 helps in providing reducing equivalents to deliver iron to Rnr2 in its reduced state but is dispensable for iron loading (Figure 4) [135]. Dre2 contains both a [2Fe-2S]+ cluster and a [4Fe-4S]2+ cluster, and participates in the CIA pathway (Figure 2) [74,137]. Like many other Fe-S assembly pathway proteins, Dre2 is essential for cell viability. The contribution of Fe-S assembly pathway components in the regulation of RNR activity was investigated by expressing genes encoding for these components under the control of the GAL1 promoter and by shifting the cells from a galactose-containing medium to a glucose-containing medium. The GAL1-mediated depletion of Dre2 reduces both the activity and the levels of Rnr2 and Rnr4 subunits in yeast [135]. The low levels of Rnr2 and Rnr4 in Dre2-depleted cells can be related to the regulation of RNR2 and RNR4 mRNA stability. In fact, Dre2 inactivation results in the increased transcription of CTH1 and CTH2 genes, which encode two RNA-binding proteins that bind specific AU-rich elements in the mRNA 3′ UTR and induce mRNA degradation [138]. The decrease in Rnr2/Rnr4 levels caused by mRNA degradation is at least in part compensated for by an induction of RNR2 and RNR4 gene transcription in the absence of Dre2. This increased transcription depends on the removal of RNR2/RNR4 transcriptional repressor Crt1 because of checkpoint activation in Dre2-depleted cells (Figure 4) [138].
The depletion of the ISC components Nfs1 and Atm1 and of the early-CIA components Cfd1, Nbp35 and Nar1 was also found to reduce RNR activity, while the depletion of the late-CIA factor Cia1 did not [139]. In addition, the depletion of Nfs1, Atm1 and Grx3/Grx4 decreases iron loading into Rnr2, while the depletion of Dre2, Cfd1, Nbp35 and Nar1 does not [139]. These results have led to the model in which iron loading into RNR requires ISC but not CIA, with the pathways for RNR and Fe-S cluster biogenesis bifurcating after Dre2-Tah18 [139]. Importantly, the maturation of Dre2 with Fe-S clusters that is necessary for Dre2 functions in vivo—including RNR regulation—depends on ISC but not on CIA [74,137]. Therefore, it appears unlikely that the late-CIA complex could modulate dNTP production through the addition of Fe-S clusters to Dre2.
Despite iron being essential for RNR activity, in both yeast and mammals, iron deficiencies result in a slight increase in the dNTP pool, suggesting that cells possess mechanisms maintaining RNR function also when iron becomes insufficient [140,141]. One of these mechanisms involves an R2 redistribution from the nucleus to the cytoplasm in response to low iron, driven by the RNA-binding proteins Cth1 and Cth2 [141]. Although the late CIA appears to be dispensable for iron radical addition to the RNR R2 subunit, it is still not known whether the CIA targeting complex contributes to regulating RNR stability or activity (Figure 3).
The depletion of either MMS19 or CIA2B results in significant changes in the intracellular levels of intermediates in nucleotide synthesis. While the amount of several metabolites is lower in the absence of a functional CIA machinery than in control cells, other central intermediates in de novo purine and pyrimidine synthesis (such as IMP and UMP) are more abundant in cells with MMS19 or CIA2B depletion than in control cells [100]. Studies that aimed to identify CIA targets revealed that CIA promotes Fe-S cluster transfer to proteins involved in nucleotide metabolisms, such as GPAT/PPAT and DPYD, which regulate de novo purine synthesis and pyrimidine catabolism, respectively [86,99,100]. Reduced DYPD protein levels were detected after the depletion of either CIA2B or MMS19 in breast cancer cells [100]. These findings suggest that CIA2B or MMS19 mutations impact nucleotide homeostasis and that an altered or unbalanced dNTP pool may contribute to the sensitivity to replication stress of cells with CIA dysfunctions. How CIA regulates dNTP production and whether and how this function supports the checkpoint-mediated regulation of dNTP availability are open questions that require further investigations.
Experiments in different organisms, spanning from E. coli and yeasts to mammalian cells, showed that either unbalanced or balanced accumulation of dNTPs dramatically increases the mutation rates during DNA replication and repair, therefore resulting in cell death or genetic abnormalities (reviewed in [33,34]). Considering that the dysregulation of the dNTP pool impacts genome stability and can sustain carcinogenesis by increasing the mutation rates and generating DNA damage, and that several approaches to treating cancer heavily exploit the effects of dNTP pool unbalance in DNA metabolism [34,35,142], the definition of the CIA-mediated control of nucleotide homeostasis can increase our understanding of mechanisms leading to cancer development and help the development of new targeted therapies.

5. Conclusions

The stability of DNA is crucial for the proper functioning of cells, while genome instability is a hallmark of cancer, which fuels carcinogenesis and the development of resistance to therapies [6]. It is increasingly clear that developing personalized therapies for the treatment of specific types of cancer is essential. This is possible if the mutations present in the tumor and their biological effects are well understood. In fact, specific mutations can make a tumor more or less sensitive to a particular type of therapy or to the use of specific inhibitors. The identification of novel potential therapeutic targets for cancer targeted therapies requires a precise understanding of how cells safeguard DNA integrity and how the involved proteins interact. In this work, we summarize the key pathways activated in response to replication stress and their regulation by Fe-S clusters, which are emerging as crucial cofactors in genome stability maintenance. However, several aspects remain unclear and require further investigation—for instance, how Fe-S clusters modulate helicase and polymerase activity during replication fork recovery, and the role of the CIA complex in regulating dNTP production.
Both replication stress and Fe-S cluster biogenesis genes have been found to be mutated in both tumors and genetic diseases [57,143,144]. In addition, iron metabolism is dysregulated in cancer cells, facilitating tumor initiation, growth and metastasis [145]. Targeting iron metabolism and modulating Fe-S clusters represent promising therapeutic strategies for cancer treatment. Current approaches include iron chelators, which restrict iron availability to tumor cells, and ferroptosis inducers, which trigger cell death through iron-dependent pathways. These therapies are being explored both as monotherapies and in combination with other cancer treatments to enhance efficacy and exploit tumor-specific vulnerabilities [144,145]. Given the interconnections between replication stress and Fe-S cluster biogenesis, the treatment of cancers harboring mutations in Fe-S cluster biogenesis genes may also benefit from the targeting of key regulators of replication stress. Indeed, it has already been observed that dysfunctions in the CIA complex sensitize triple-negative breast cancer cells to ATR and CHK1 inhibitors [100]. Identifying additional proteins involved in replication stress and Fe-S cluster biogenesis, as well as exploring their interconnections, could open new avenues for targeted therapies.

Author Contributions

Conceptualization, C.F., M.G. and M.C.; writing—original draft preparation, C.F., M.G., S.C., A.D.P. and M.C.; writing—review and editing, C.F., M.G., S.C., A.D.P. and M.C.; funding acquisition, M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the European Union—Next Generation EU, Mission 4 Component 2—CUP H53D23004830006—INVESTIMENTO 1.1, Fondo per il Programma Nazionale di Ricerca e Progetti di Rilevante Interesse Nazionale (PRIN2022) del Piano Nazionale di Ripresa e Resilienza.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank all members of the labs for helpful discussion.

Conflicts of Interest

The authors declare no conflicts 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]
  2. Saxena, S.; Zou, L. Hallmarks of DNA replication stress. Mol. Cell 2022, 82, 2298–2314. [Google Scholar] [CrossRef] [PubMed]
  3. Giannattasio, M.; Branzei, D. S-phase checkpoint regulations that preserve replication and chromosome integrity upon dNTP depletion. Cell. Mol. Life Sci. 2017, 74, 2361–2380. [Google Scholar] [CrossRef] [PubMed]
  4. Pasero, P.; Vindigni, A. Nucleases acting at stalled forks: How to reboot the replication program with a few shortcuts. Annu. Rev. Genet. 2017, 51, 477–499. [Google Scholar] [CrossRef]
  5. Gaillard, H.; García-Muse, T.; Aguilera, A. Replication stress and cancer. Nat. Rev. Cancer 2015, 15, 276–289. [Google Scholar] [CrossRef]
  6. Hanahan, D. Hallmarks of cancer: New dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef] [PubMed]
  7. Paul, V.D.; Lill, R. Biogenesis of cytosolic and nuclear iron–sulfur proteins and their role in genome stability. Biochim. Biophys. Acta Mol. Cell Res. 2015, 1853, 1528–1539. [Google Scholar] [CrossRef]
  8. Baranovskiy, A.G.; Duong, V.N.; Babayeva, N.D.; Zhang, Y.; Pavlov, Y.I.; Anderson, K.S.; Tahirov, T.H. Activity and fidelity of human DNA polymerase α depend on primer structure. J. Biol. Chem. 2018, 293, 6824–6843. [Google Scholar] [CrossRef]
  9. Burgers, P.M.J.; Kunkel, T.A. Eukaryotic DNA Replication Fork. Annu. Rev. Biochem. 2017, 86, 417–438. [Google Scholar] [CrossRef]
  10. Gambus, A.; Van Deursen, F.; Polychronopoulos, D.; Foltman, M.; Jones, R.C.; Edmondson, R.D.; Calzada, A.; Labib, K. A key role for Ctf4 in coupling the MCM2-7 helicase to DNA polymerase α within the eukaryotic replisome. EMBO J. 2009, 28, 2992–3004. [Google Scholar] [CrossRef]
  11. Kang, Y.H.; Farina, A.; Bermudez, V.P.; Tappin, I.; Du, F.; Galal, W.C.; Hurwitz, J. Interaction between human Ctf4 and the Cdc45/Mcm2-7/GINS (CMG) replicative helicase. Proc. Natl. Acad. Sci. USA 2013, 110, 19760–19765. [Google Scholar] [CrossRef] [PubMed]
  12. Simon, A.C.; Zhou, J.C.; Perera, R.L.; van Deursen, F.; Evrin, C.; Ivanova, M.E.; Kilkenny, M.L.; Renault, L.; Kjaer, S.; Matak-Vinković, D.; et al. A Ctf4 trimer couples the CMG helicase to DNA polymerase α in the eukaryotic replisome. Nature 2014, 510, 293–297. [Google Scholar] [CrossRef]
  13. Villa, F.; Simon, A.C.; Bazan, M.A.O.; Kilkenny, M.L.; Wirthensohn, D.; Wightman, M.; Matak-Vinković, D.; Pellegrini, L.; Labib, K. Ctf4 is a hub in the eukaryotic replisome that links multiple CIP-box proteins to the CMG helicase. Mol. Cell 2016, 63, 385–396. [Google Scholar] [CrossRef] [PubMed]
  14. Kokoska, R.J.; Stefanovic, L.; DeMai, J.; Petes, T.D. Increased rates of genomic deletions generated by mutations in the yeast gene encoding DNA polymerase δ or by decreases in the cellular levels of DNA polymerase δ. Mol. Cell. Biol. 2000, 20, 362–373. [Google Scholar] [CrossRef]
  15. Netz, D.J.; Stith, C.M.; Stümpfig, M.; Köpf, G.; Vogel, D.; Genau, H.M.; Stodola, J.L.; Lill, R.; Burgers, P.M.; Pierik, A.J. Eukaryotic DNA polymerases require an iron-sulfur cluster for the formation of active complexes. Nat. Chem. Biol. 2011, 8, 125–132. [Google Scholar] [CrossRef] [PubMed]
  16. Morrison, A.; Araki, H.; Clark, A.B.; Hamatake, R.K.; Sugino, A. A third essential DNA polymerase in S. cerevisiae. Cell 1990, 62, 1143–1151. [Google Scholar] [CrossRef]
  17. Burgers, P.M. Eukaryotic DNA polymerases in DNA replication and DNA repair. Chromosoma 1998, 107, 218–227. [Google Scholar] [CrossRef]
  18. Zhou, J.C.; Janska, A.; Goswami, P.; Renault, L.; Abid Ali, F.; Kotecha, A.; Diffley, J.F.X.; Costa, A. CMG–Pol epsilon dynamics suggests a mechanism for the establishment of leading-strand synthesis in the eukaryotic replisome. Proc. Natl. Acad. Sci. USA 2017, 114, 4141–4146. [Google Scholar] [CrossRef]
  19. Hogg, M.; Johansson, E. DNA polymerase ε. Eukaryot. Replisome Guide Protein Struct. Funct. 2012, 62, 237–257. [Google Scholar]
  20. Klinge, S.; Hirst, J.; Maman, J.D.; Krude, T.; Pellegrini, L. An iron-sulfur domain of the eukaryotic primase is essential for RNA primer synthesis. Nat. Struct. Mol. Biol. 2007, 14, 875–877. [Google Scholar] [CrossRef]
  21. Weiner, B.E.; Huang, H.; Dattilo, B.M.; Nilges, M.J.; Fanning, E.; Chazin, W.J. An iron-sulfur cluster in the C-terminal domain of the p58 subunit of human DNA primase. J. Biol. Chem. 2007, 282, 33444–33451. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, P.; Carvalho, C.M.; Hastings, P.J.; Lupski, J.R. Mechanisms for recurrent and complex human genomic rearrangements. Curr. Opin. Genet. Dev. 2012, 22, 211–220. [Google Scholar] [CrossRef] [PubMed]
  23. Segurado, M.; Tercero, J.A. The S-phase checkpoint: Targeting the replication fork. Biol. Cell 2009, 101, 617–627. [Google Scholar] [CrossRef]
  24. Berti, M.; Cortez, D.; Lopes, M. The plasticity of DNA replication forks in response to clinically relevant genotoxic stress. Nat. Rev. Mol. Cell Biol. 2020, 21, 633–651. [Google Scholar] [CrossRef]
  25. Dueva, R.; Iliakis, G. Replication protein A: A multifunctional protein with roles in DNA replication, repair and beyond. NAR Cancer 2020, 2, zcaa022. [Google Scholar] [CrossRef]
  26. Hashimoto, Y.; Ray Chaudhuri, A.; Lopes, M.; Costanzo, V. Rad51 protects nascent DNA from Mre11-dependent degradation and promotes continuous DNA synthesis. Nat. Struct. Mol. Biol. 2010, 17, 1305–1311. [Google Scholar] [CrossRef]
  27. Pizzul, P.; Casari, E.; Gnugnoli, M.; Rinaldi, C.; Corallo, F.; Longhese, M.P. The DNA damage checkpoint: A tale from budding yeast. Front. Genet. 2022, 13, 995163. [Google Scholar] [CrossRef] [PubMed]
  28. Tercero, J.A.; Longhese, M.P.; Diffley, J.F. A central role for DNA replication forks in checkpoint activation and response. Mol. Cell 2003, 11, 1323–1336. [Google Scholar] [CrossRef]
  29. Pardo, B.; Crabbé, L.; Pasero, P. Signaling pathways of replication stress in yeast. FEMS Yeast Res. 2017, 17, fow101. [Google Scholar] [CrossRef]
  30. Bando, M.; Katou, Y.; Komata, M.; Tanaka, H.; Itoh, T.; Sutani, T.; Shirahige, K. Csm3, Tof1, and Mrc1 form a heterotrimeric mediator complex that associates with DNA replication forks. J. Biol. Chem. 2009, 284, 34355–34365. [Google Scholar] [CrossRef]
  31. Ciardo, D.; Goldar, A.; Marheineke, K. On the interplay of the DNA replication program and the intra-S phase checkpoint pathway. Genes 2019, 10, 94. [Google Scholar] [CrossRef] [PubMed]
  32. Zegerman, P.; Diffley, J.F. Checkpoint-dependent inhibition of DNA replication initiation by Sld3 and Dbf4 phosphorylation. Nature 2010, 467, 474–478. [Google Scholar] [CrossRef] [PubMed]
  33. Sanvisens, N.; de Llanos, R.; Puig, S. Function and regulation of yeast ribonucleotide reductase: Cell cycle, genotoxic stress, and iron bioavailability. Biomed. J. 2013, 36, 51–58. [Google Scholar]
  34. Mathews, C.K. Deoxyribonucleotide metabolism, mutagenesis and cancer. Nat. Rev. Cancer 2015, 15, 528–539. [Google Scholar] [CrossRef] [PubMed]
  35. Yagüe-Capilla, M.; Rudd, S.G. Understanding the interplay between dNTP metabolism and genome stability in cancer. Dis. Models Mech. 2024, 17, dmm050775. [Google Scholar] [CrossRef] [PubMed]
  36. Elledge, S.J.; Davis, R.W. Two genes differentially regulated in the cell cycle and by DNA-damaging agents encode alternative regulatory subunits of ribonucleotide reductase. Genes Dev. 1990, 4, 740–751. [Google Scholar] [CrossRef]
  37. Domkin, V.; Thelander, L.; Chabes, A. Yeast DNA damage-inducible Rnr3 has a very low catalytic activity strongly stimulated after the formation of a cross-talking Rnr1/Rnr3 complex. J. Biol. Chem. 2002, 277, 18574–18578. [Google Scholar] [CrossRef]
  38. Huang, M.; Zhou, Z.; Elledge, S.J. The DNA replication and damage checkpoint pathways induce transcription by inhibition of the Crt1 repressor. Cell 1998, 94, 595–605. [Google Scholar] [CrossRef]
  39. Guarino, E.; Salguero, I.; Kearsey, S.E. Cellular regulation of ribonucleotide reductase in eukaryotes. Semin. Cell Dev. Biol. 2014, 30, 97–103. [Google Scholar] [CrossRef]
  40. D’Angiolella, V.; Donato, V.; Forrester, F.M.; Jeong, Y.T.; Pellacani, C.; Kudo, Y.; Saraf, A.; Florens, L.; Washburn, M.P.; Pagano, M. Cyclin F-mediated degradation of ribonucleotide reductase M2 controls genome integrity and DNA repair. Cell 2012, 149, 1023–1034. [Google Scholar] [CrossRef]
  41. Puig, S.; Ramos-Alonso, L.; Romero, A.M.; Martínez-Pastor, M.T. The elemental role of iron in DNA synthesis and repair. Metallomics 2017, 9, 1483–1500. [Google Scholar] [CrossRef]
  42. Ge, X.Q.; Blow, J.J. Chk1 inhibits replication factory activation but allows dormant origin firing in existing factories. J. Cell Biol. 2010, 191, 1285–1297. [Google Scholar] [CrossRef] [PubMed]
  43. Branzei, D.; Psakhye, I. DNA damage tolerance. Curr. Opin. Cell Biol. 2016, 40, 137–144. [Google Scholar] [CrossRef] [PubMed]
  44. Bianchi, J.; Rudd, S.G.; Jozwiakowski, S.K.; Bailey, L.J.; Soura, V.; Taylor, E.; Stevanovic, I.; Green, A.J.; Stracker, T.H.; Lindsay, H.D.; et al. PrimPol bypasses UV photoproducts during eukaryotic chromosomal DNA replication. Mol. Cell 2013, 52, 566–573. [Google Scholar] [CrossRef] [PubMed]
  45. Sale, J.E.; Lehmann, A.R.; Woodgate, R. Y-family DNA polymerases and their role in tolerance of cellular DNA damage. Nat. Rev. Mol. Cell Biol. 2012, 13, 141–152. [Google Scholar] [CrossRef]
  46. Atkinson, J.; McGlynn, P. Replication fork reversal and the maintenance of genome stability. Nucleic Acids Res. 2009, 37, 3475–3492. [Google Scholar] [CrossRef]
  47. Sogo, J.M.; Lopes, M.; Foiani, M. Fork reversal and ssDNA accumulation at stalled replication forks owing to checkpoint defects. Science 2002, 297, 599–602. [Google Scholar] [CrossRef]
  48. 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]
  49. Berti, M.; Vindigni, A. Replication stress: Getting back on track. Nat. Struct. Mol. Biol. 2016, 23, 103–109. [Google Scholar] [CrossRef]
  50. Yeeles, J.T.; Janska, A.; Early, A.; Diffley, J.F. How the eukaryotic replisome achieves rapid and efficient DNA replication. Mol. Cell 2017, 65, 105–116. [Google Scholar] [CrossRef]
  51. Petronek, M.S.; Allen, B.G. Maintenance of genome integrity by the late-acting cytoplasmic iron-sulfur assembly (CIA) complex. Front. Genet. 2023, 14, 1152398. [Google Scholar] [CrossRef]
  52. Lill, R.; Mühlenhoff, U. Iron-sulfur protein biogenesis in eukaryotes: Components and mechanisms. Annu. Rev. Cell Dev. Biol. 2006, 22, 457–486. [Google Scholar] [CrossRef] [PubMed]
  53. Lill, R.; Dutkiewicz, R.; Elsässer, H.P.; Hausmann, A.; Netz, D.J.; Pierik, A.J.; Stehling, O.; Urzica, E.; Mühlenhoff, U. Mechanisms of iron–sulfur protein maturation in mitochondria, cytosol, and nucleus of eukaryotes. Biochim. Biophys. Acta Mol. Cell Res. 2006, 1763, 652–667. [Google Scholar] [CrossRef] [PubMed]
  54. Rydz, L.; Wróbel, M.; Jurkowska, H. Sulfur administration in Fe–S cluster homeostasis. Antioxidants 2021, 10, 1738. [Google Scholar] [CrossRef] [PubMed]
  55. Shi, R.; Hou, W.; Wang, Z.Q.; Xu, X. Biogenesis of iron–sulfur clusters and their role in DNA metabolism. Front. Cell Dev. Biol. 2021, 9, 735678. [Google Scholar] [CrossRef]
  56. White, M.F.; Dillingham, M.S. Iron–sulfur clusters in nucleic acid processing enzymes. Curr. Opin. Struct. Biol. 2012, 22, 94–100. [Google Scholar] [CrossRef]
  57. Maio, N.; Rouault, T.A. Iron–sulfur cluster biogenesis in mammalian cells: New insights into the molecular mechanisms of cluster delivery. Biochim. Biophys. Acta Mol. Cell Res. 2015, 1853, 1493–1512. [Google Scholar] [CrossRef]
  58. Valasatava, Y.; Rosato, A.; Banci, L.; Andreini, C. MetalPredator: A web server to predict iron-sulfur cluster binding proteomes. Bioinformatics 2016, 32, 2850–2852. [Google Scholar] [CrossRef]
  59. Wehrspan, Z.J.; McDonnell, R.T.; Elcock, A.H. Identification of iron-sulfur (Fe-S) cluster and zinc (Zn) binding sites within proteomes predicted by DeepMind’s AlphaFold2 program dramatically expands the metalloproteome. J. Mol. Biol. 2022, 434, 167377. [Google Scholar] [CrossRef]
  60. Vallières, C.; Benoit, O.; Guittet, O.; Huang, M.E.; Lepoivre, M.; Golinelli-Cohen, M.P.; Vernis, L. Iron-sulfur protein odyssey: Exploring their cluster functional versatility and challenging identification. Metallomics 2024, 16, mfae025. [Google Scholar] [CrossRef]
  61. Netz, D.J.; Mascarenhas, J.; Stehling, O.; Pierik, A.J.; Lill, R. Maturation of cytosolic and nuclear iron–sulfur proteins. Trends Cell Biol. 2014, 24, 303–312. [Google Scholar] [CrossRef] [PubMed]
  62. Ayala-Castro, C.; Saini, A.; Outten, F.W. Fe-S cluster assembly pathways in bacteria. Microbiol. Mol. Biol. Rev. 2008, 72, 110–125. [Google Scholar] [CrossRef] [PubMed]
  63. Mettert, E.L.; Kiley, P.J. Fe–S proteins that regulate gene expression. Biochim. Biophys. Acta Mol. Cell Res. 2015, 1853, 1284–1293. [Google Scholar] [CrossRef]
  64. Wu, Y.; Brosh, R.M., Jr. DNA helicase and helicase–nuclease enzymes with a conserved iron–sulfur cluster. Nucleic Acids Res. 2012, 40, 4247–4260. [Google Scholar] [CrossRef]
  65. Adrover, M.; Howes, B.D.; Iannuzzi, C.; Smulevich, G.; Pastore, A. Anatomy of an iron-sulfur cluster scaffold protein: Understanding the determinants of [2Fe–2S] cluster stability on IscU. Biochim. Biophys. Acta (BBA)-Mol. Cell Res. 2015, 1853, 1448–1456. [Google Scholar] [CrossRef] [PubMed]
  66. Uzarska, M.A.; Grochowina, I.; Soldek, J.; Jelen, M.; Schilke, B.; Marszalek, J.; Craig, E.A.; Dutkiewicz, R. During FeS cluster biogenesis, ferredoxin and frataxin use overlapping binding sites on yeast cysteine desulfurase Nfs1. J. Biol. Chem. 2022, 298, 101570. [Google Scholar] [CrossRef]
  67. Webert, H.; Freibert, S.A.; Gallo, A.; Heidenreich, T.; Linne, U.; Amlacher, S.; Hurt, E.; Mühlenhoff, U.; Banci, L.; Lill, R. Functional reconstitution of mitochondrial Fe/S cluster synthesis on Isu1 reveals the involvement of ferredoxin. Nat. Commun. 2014, 5, 5013. [Google Scholar] [CrossRef]
  68. Galeano, B.K.; Ranatunga, W.; Gakh, O.; Smith, D.Y., IV; Thompson, J.R.; Isaya, G. Zinc and the iron donor frataxin regulate oligomerization of the scaffold protein to form new Fe–S cluster assembly centers. Metallomics 2017, 9, 773–801. [Google Scholar] [CrossRef]
  69. Ciofi-Baffoni, S.; Nasta, V.; Banci, L. Protein networks in the maturation of human iron–sulfur proteins. Metallomics 2018, 10, 49–72. [Google Scholar] [CrossRef]
  70. Voisine, C.; Cheng, Y.C.; Ohlson, M.; Schilke, B.; Hoff, K.; Beinert, H.; Marszalek, J.; Craig, E.A. Jac1, a mitochondrial J-type chaperone, is involved in the biogenesis of Fe/S clusters in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 2001, 98, 1483–1488. [Google Scholar] [CrossRef]
  71. Melber, A.; Na, U.; Vashisht, A.; Weiler, B.D.; Lill, R.; Wohlschlegel, J.A.; Winge, D.R. Role of Nfu1 and Bol3 in iron-sulfur cluster transfer to mitochondrial clients. Elife 2016, 5, e15991. [Google Scholar] [CrossRef]
  72. Li, P.; Hendricks, A.L.; Wang, Y.; Villones, R.L.E.; Lindkvist-Petersson, K.; Meloni, G.; Cowan, J.A.; Wang, K.; Gourdon, P. Structures of Atm1 provide insight into [2Fe-2S] cluster export from mitochondria. Nat. Commun. 2022, 13, 4339. [Google Scholar] [CrossRef] [PubMed]
  73. Netz, D.J.; Pierik, A.J.; Stümpfig, M.; Bill, E.; Sharma, A.K.; Pallesen, L.J.; Walden, W.E.; Lill, R. A bridging [4Fe-4S] cluster and nucleotide binding are essential for function of the Cfd1-Nbp35 complex as a scaffold in iron-sulfur protein maturation. J. Biol. Chem. 2012, 287, 12365–12378. [Google Scholar] [CrossRef] [PubMed]
  74. Netz, D.J.; Stümpfig, M.; Doré, C.; Mühlenhoff, U.; Pierik, A.J.; Lill, R. Tah18 transfers electrons to Dre2 in cytosolic iron-sulfur protein biogenesis. Nat. Chem. Biol. 2010, 6, 758–765. [Google Scholar] [CrossRef] [PubMed]
  75. Stehling, O.; Lill, R. The role of mitochondria in cellular iron–sulfur protein biogenesis: Mechanisms, connected processes, and diseases. Cold Spring Harb. Perspect. Biol. 2013, 5, a011312. [Google Scholar]
  76. Mühlenhoff, U.; Braymer, J.J.; Christ, S.; Rietzschel, N.; Uzarska, M.A.; Weiler, B.D.; Lill, R. Glutaredoxins and iron-sulfur protein biogenesis at the interface of redox biology and iron metabolism. Biol. Chem. 2020, 401, 1407–1428. [Google Scholar] [CrossRef]
  77. Lill, R. From the discovery to molecular understanding of cellular iron-sulfur protein biogenesis. Biol. Chem. 2020, 401, 855–876. [Google Scholar] [CrossRef]
  78. Lill, R.; Freibert, S.A. Mechanisms of mitochondrial iron-sulfur protein biogenesis. Annu. Rev. Biochem. 2020, 89, 471–499. [Google Scholar] [CrossRef]
  79. Vo, A.T.; Fleischman, N.M.; Marquez, M.D.; Camire, E.J.; Esonwune, S.U.; Grossman, J.D.; Gay, K.A.; Cosman, J.A.; Perlstein, D.L. Defining the domains of Cia2 required for its essential function in vivo and in vitro. Metallomics 2017, 9, 1645–1654. [Google Scholar] [CrossRef]
  80. Srinivasan, V.; Netz, D.J.; Webert, H.; Mascarenhas, J.; Pierik, A.J.; Michel, H.; Lill, R. Structure of the yeast WD40 domain protein Cia1, a component acting late in iron-sulfur protein biogenesis. Structure 2007, 15, 1246–1257. [Google Scholar] [CrossRef]
  81. Vo, A.; Fleischman, N.M.; Froehlich, M.J.; Lee, C.Y.; Cosman, J.A.; Glynn, C.A.; Hassan, Z.O.; Perlstein, D.L. Identifying the protein interactions of the cytosolic iron–sulfur cluster targeting complex essential for its assembly and recognition of apo-targets. Biochemistry 2018, 57, 2349–2358. [Google Scholar] [CrossRef] [PubMed]
  82. Johnstone, R.W.; Wang, J.; Tommerup, N.; Vissing, H.; Roberts, T.; Shi, Y. Ciao 1 is a novel WD40 protein that interacts with the tumor suppressor protein WT1. J. Biol. Chem. 1998, 273, 10880–10887. [Google Scholar] [CrossRef]
  83. Balk, J.; Aguilar Netz, D.J.; Tepper, K.; Pierik, A.J.; Lill, R. The essential WD40 protein Cia1 is involved in a late step of cytosolic and nuclear iron-sulfur protein assembly. Mol. Cell. Biol. 2005, 25, 10833–10841. [Google Scholar] [CrossRef] [PubMed]
  84. Maio, N.; Orbach, R.; Zaharieva, I.T.; Töpf, A.; Donkervoort, S.; Munot, P.; Mueller, J.; Willis, T.; Verma, S.; Peric, S.; et al. CIAO1 loss of function causes a neuromuscular disorder with compromise of nucleocytoplasmic Fe-S enzymes. J. Clin. Investig. 2024, 134, e179559. [Google Scholar] [CrossRef] [PubMed]
  85. Braymer, J.J.; Stehling, O.; Stümpfig, M.; Rösser, R.; Spantgar, F.; Blinn, C.M.; Mühlenhoff, U.; Pierik, A.J.; Lill, R. Requirements for the biogenesis of [2Fe-2S] proteins in the human and yeast cytosol. Proc. Natl. Acad. Sci. USA 2024, 121, e2400740121. [Google Scholar] [CrossRef]
  86. Stehling, O.; Mascarenhas, J.; Vashisht, A.A.; Sheftel, A.D.; Niggemeyer, B.; Rösser, R.; Pierik, A.J.; Wohlschlegel, J.A.; Lill, R. Human CIA2A-FAM96A and CIA2B-FAM96B integrate iron homeostasis and maturation of different subsets of cytosolic-nuclear iron-sulfur proteins. Cell Metab. 2013, 18, 187–198. [Google Scholar] [CrossRef]
  87. Maione, V.; Cantini, F.; Severi, M.; Banci, L. Investigating the role of the human CIA2A-CIAO1 complex in the maturation of aconitase. Biochim. Biophys. Acta Gen. Subj. 2018, 1862, 1980–1987. [Google Scholar] [CrossRef]
  88. Ito, S.; Tan, L.J.; Andoh, D.; Narita, T.; Seki, M.; Hirano, Y.; Narita, K.; Kuraoka, I.; Hiraoka, Y.; Tanaka, K. MMXD, a TFIIH-independent XPD-MMS19 protein complex involved in chromosome segregation. Mol. Cell 2010, 39, 632–640. [Google Scholar] [CrossRef]
  89. Zhang, X.H.; Chen, X.J.; Yan, Y.B. Fam96b recruits brain-type creatine kinase to fuel mitotic spindle formation. Biochim. Biophys. Acta (BBA)-Mol. Cell Res. 2023, 1870, 119410. [Google Scholar] [CrossRef]
  90. Lev, I.; Volpe, M.; Goor, L.; Levinton, N.; Emuna, L.; Ben-Aroya, S. Reverse PCA, a systematic approach for identifying genes important for the physical interaction between protein pairs. PLoS Genet. 2013, 9, e1003838. [Google Scholar] [CrossRef]
  91. Weerapana, E.; Wang, C.; Simon, G.M.; Richter, F.; Khare, S.; Dillon, M.B.; Bachovchin, D.A.; Mowen, K.; Baker, D.; Cravatt, B.F. Quantitative reactivity profiling predicts functional cysteines in proteomes. Nature 2010, 468, 790–795. [Google Scholar] [CrossRef] [PubMed]
  92. Gari, K.; León Ortiz, A.M.; Borel, V.; Flynn, H.; Skehel, J.M.; Boulton, S.J. MMS19 links cytoplasmic iron-sulfur cluster assembly to DNA metabolism. Science 2012, 337, 243–245. [Google Scholar] [CrossRef] [PubMed]
  93. Kassube, S.A.; Thomä, N.H. Structural insights into Fe–S protein biogenesis by the CIA targeting complex. Nat. Struct. Mol. Biol. 2020, 27, 735–742. [Google Scholar] [CrossRef]
  94. Vasquez, S.; Marquez, M.D.; Brignole, E.J.; Vo, A.; Kong, S.; Park, C.; Perlstein, D.L.; Drennan, C.L. Structural and biochemical investigations of a HEAT-repeat protein involved in the cytosolic iron-sulfur cluster assembly pathway. Commun. Biol. 2023, 6, 1276. [Google Scholar] [CrossRef] [PubMed]
  95. Seki, M.; Takeda, Y.; Iwai, K.; Tanaka, K. IOP1 protein is an external component of the human cytosolic iron-sulfur cluster assembly (CIA) machinery and functions in the MMS19 protein-dependent CIA pathway. J. Biol. Chem. 2013, 288, 16680–16689. [Google Scholar] [CrossRef]
  96. Prakash, L.; Prakash, S. Three additional genes involved in pyrimidine dimer removal in Saccharomyces cerevisiae: RAD7, RAD14, and MMS19. Mol. Gen. Genet. (MGG) 1979, 176, 351–359. [Google Scholar] [CrossRef]
  97. Chang, M.; Bellaoui, M.; Boone, C.; Brown, G.W. A genome-wide screen for methyl methanesulfonate-sensitive mutants reveals genes required for S phase progression in the presence of DNA damage. Proc. Natl. Acad. Sci. USA 2002, 99, 16934–16939. [Google Scholar] [CrossRef]
  98. Kou, H.; Zhou, Y.; Gorospe, R.C.; Wang, Z. Mms19 protein functions in nucleotide excision repair by sustaining an adequate cellular concentration of the TFIIH component Rad3. Proc. Natl. Acad. Sci. USA 2008, 105, 15714–15719. [Google Scholar] [CrossRef]
  99. Stehling, O.; Vashisht, A.A.; Mascarenhas, J.; Jonsson, Z.O.; Sharma, T.; Netz, D.J.; Pierik, A.J.; Wohlschlegel, J.A.; Lill, R. MMS19 assembles iron-sulfur proteins required for DNA metabolism and genomic integrity. Science 2012, 337, 195–199. [Google Scholar] [CrossRef]
  100. Redwood, A.B.; Zhang, X.; Seth, S.B.; Ge, Z.; Bindeman, W.E.; Zhou, X.; Sinha, V.C.; Heffernan, T.P.; Piwnica-Worms, H. The cytosolic iron–sulfur cluster assembly (CIA) pathway is required for replication stress tolerance of cancer cells to Chk1 and ATR inhibitors. NPJ Breast Cancer 2021, 7, 152. [Google Scholar] [CrossRef]
  101. Corcoles-Saez, I.; Dong, K.; Johnson, A.L.; Waskiewicz, E.; Costanzo, M.; Boone, C.; Cha, R.S. Essential function of Mec1, the budding yeast ATM/ATR checkpoint-response kinase, in protein homeostasis. Dev. Cell 2018, 46, 495–503. [Google Scholar] [CrossRef]
  102. de Oliveira, F.M.B.; Kim, D.; Cussiol, J.R.; Das, J.; Jeong, M.C.; Doerfler, L.; Schmidt, K.H.; Yu, H.; Smolka, M.B. Phosphoproteomics reveals distinct modes of Mec1/ATR signaling during DNA replication. Mol. Cell 2015, 57, 1124–1132. [Google Scholar] [CrossRef]
  103. Lanz, M.C.; Oberly, S.; Sanford, E.J.; Sharma, S.; Chabes, A.; Smolka, M.B. Separable roles for Mec1/ATR in genome maintenance, DNA replication, and checkpoint signaling. Genes Dev. 2018, 32, 822–835. [Google Scholar] [CrossRef] [PubMed]
  104. Forey, R.; Poveda, A.; Sharma, S.; Barthe, A.; Padioleau, I.; Renard, C.; Lambert, R.; Skrzypczak, M.; Ginalski, K.; Lengronne, A.; et al. Mec1 is activated at the onset of normal S phase by low-dNTP pools impeding DNA replication. Mol. Cell 2020, 78, 396–410. [Google Scholar] [CrossRef]
  105. Giot, L.; Chanet, R.; Simon, M.; Facca, C.; Faye, G. Involvement of the yeast DNA polymerase δ in DNA repair in vivo. Genetics 1997, 146, 1239–1251. [Google Scholar] [CrossRef] [PubMed]
  106. Baranovskiy, A.G.; Lada, A.G.; Siebler, H.M.; Zhang, Y.; Pavlov, Y.I.; Tahirov, T.H. DNA polymerase δ and ζ switch by sharing accessory subunits of DNA polymerase δ. J. Biol. Chem. 2012, 287, 17281–17287. [Google Scholar] [CrossRef] [PubMed]
  107. Baranovskiy, A.G.; Siebler, H.M.; Pavlov, Y.I.; Tahirov, T.H. Iron–sulfur clusters in DNA polymerases and primases of eukaryotes. Methods Enzymol. 2018, 599, 1–20. [Google Scholar]
  108. Jain, R.; Rice, W.J.; Malik, R.; Johnson, R.E.; Prakash, L.; Prakash, S.; Ubarretxena-Belandia, I.; Aggarwal, A.K. Cryo-EM structure and dynamics of eukaryotic DNA polymerase δ holoenzyme. Nat. Struct. Mol. Biol. 2019, 26, 955–962. [Google Scholar] [CrossRef]
  109. Jain, R.; Vanamee, E.S.; Dzikovski, B.G.; Buku, A.; Johnson, R.E.; Prakash, L.; Prakash, S.; Aggarwal, A.K. An iron–sulfur cluster in the polymerase domain of yeast DNA polymerase ε. J. Mol. Biol. 2014, 426, 301–308. [Google Scholar] [CrossRef]
  110. Ter Beek, J.; Parkash, V.; Bylund, G.O.; Osterman, P.; Sauer-Eriksson, A.E.; Johansson, E. Structural evidence for an essential Fe–S cluster in the catalytic core domain of DNA polymerase ϵ. Nucleic Acids Res. 2019, 47, 5712–5722. [Google Scholar] [CrossRef]
  111. Lisova, A.E.; Baranovskiy, A.G.; Morstadt, L.M.; Babayeva, N.D.; Stepchenkova, E.I.; Tahirov, T.H. The iron-sulfur cluster is essential for DNA binding by human DNA polymerase ε. Sci. Rep. 2022, 12, 17436. [Google Scholar] [CrossRef]
  112. Barton, J.K.; Silva, R.M.; O’Brien, E. Redox chemistry in the genome: Emergence of the [4Fe4S] cofactor in repair and replication. Annu. Rev. Biochem. 2019, 88, 163–190. [Google Scholar] [CrossRef] [PubMed]
  113. O’Brien, E.; Holt, M.E.; Thompson, M.K.; Salay, L.E.; Ehlinger, A.C.; Chazin, W.J.; Barton, J.K. The [4Fe4S] cluster of human DNA primase functions as a redox switch using DNA charge transport. Science 2017, 355, eaag1789. [Google Scholar] [CrossRef]
  114. Bartels, P.L.; Stodola, J.L.; Burgers, P.M.; Barton, J.K. A redox role for the [4Fe4S] cluster of yeast DNA polymerase δ. J. Am. Chem. Soc. 2017, 139, 18339–18348. [Google Scholar] [CrossRef] [PubMed]
  115. Pinto, M.N.; Ter Beek, J.; Ekanger, L.A.; Johansson, E.; Barton, J.K. The [4Fe4S] cluster of yeast DNA polymerase ε is redox active and can undergo DNA-mediated signaling. J. Am. Chem. Soc. 2021, 143, 16147–16153. [Google Scholar] [CrossRef]
  116. 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]
  117. Shaw, A.E.; Whitted, J.E.; Mihelich, M.N.; Reitman, H.J.; Timmerman, A.J.; Schauer, G.D. Revised Mechanism of Hydroxyurea Induced Cell Cycle Arrest and an Improved Alternative. Proc. Natl. Acad. Sci. USA 2024, 121, e2404470121. [Google Scholar] [CrossRef]
  118. Jozwiakowski, S.K.; Kummer, S.; Gari, K. Human DNA polymerase delta requires an iron–sulfur cluster for high-fidelity DNA synthesis. Life Sci. Alliance 2019, 2, e201900321. [Google Scholar] [CrossRef]
  119. Chanet, R.; Baïlle, D.; Golinelli-Cohen, M.P.; Riquier, S.; Guittet, O.; Lepoivre, M.; Huang, M.E.; Vernis, L. Fe-S coordination defects in the replicative DNA polymerase delta cause deleterious DNA replication in vivo and subsequent DNA damage in the yeast Saccharomyces cerevisiae. G3 2021, 11, jkab124. [Google Scholar] [CrossRef]
  120. Kiktev, D.A.; Dominska, M.; Zhang, T.; Dahl, J.; Stepchenkova, E.I.; Mieczkowski, P.; Burgers, P.M.; Lujan, S.; Burkholder, A.; Kunkel, T.A.; et al. The fidelity of DNA replication, particularly on GC-rich templates, is reduced by defects of the Fe–S cluster in DNA polymerase δ. Nucleic Acids Res. 2021, 49, 5623–5636. [Google Scholar] [CrossRef]
  121. Fuss, J.O.; Tsai, C.L.; Ishida, J.P.; Tainer, J.A. Emerging critical roles of Fe-S clusters in DNA replication and repair. Biochim. Biophys. Acta 2015, 1853, 1253–1271. [Google Scholar] [CrossRef]
  122. Rossi, S.E.; Foiani, M.; Giannattasio, M. Dna2 processes behind the fork long ssDNA flaps generated by Pif1 and replication-dependent strand displacement. Nat. Commun. 2018, 9, 4830. [Google Scholar] [CrossRef]
  123. Hu, J.; Sun, L.; Shen, F.; Chen, Y.; Hua, Y.; Liu, Y.; Zhang, M.; Hu, Y.; Wang, Q.; Xu, W.; et al. The intra-S phase checkpoint targets Dna2 to prevent stalled replication forks from reversing. Cell 2012, 149, 1221–1232. [Google Scholar] [CrossRef] [PubMed]
  124. Thangavel, S.; Berti, M.; Levikova, M.; Pinto, C.; Gomathinayagam, S.; Vujanovic, M.; Zellweger, R.; Moore, H.; Lee, E.H.; Hendrickson, E.A.; et al. DNA2 drives processing and restart of reversed replication forks in human cells. J. Cell Biol. 2015, 208, 545–562. [Google Scholar] [CrossRef]
  125. Yeeles, J.T.; Cammack, R.; Dillingham, M.S. An iron-sulfur cluster is essential for the binding of broken DNA by AddAB-type helicase-nucleases. J. Biol. Chem. 2009, 284, 7746–7755. [Google Scholar] [CrossRef] [PubMed]
  126. Pokharel, S.; Campbell, J.L. Cross talk between the nuclease and helicase activities of Dna2: Role of an essential iron–sulfur cluster domain. Nucleic Acids Res. 2012, 40, 7821–7830. [Google Scholar] [CrossRef]
  127. Zhou, C.; Pourmal, S.; Pavletich, N.P. Dna2 nuclease-helicase structure, mechanism and regulation by Rpa. Elife 2015, 4, e09832. [Google Scholar] [CrossRef] [PubMed]
  128. Mariotti, L.; Wild, S.; Brunoldi, G.; Piceni, A.; Ceppi, I.; Kummer, S.; Lutz, R.E.; Cejka, P.; Gari, K. The iron–sulfur cluster in human DNA2 is required for all biochemical activities of DNA2. Commun. Biol. 2020, 3, 322. [Google Scholar] [CrossRef]
  129. van der Lelij, P.; Chrzanowska, K.H.; Godthelp, B.C.; Rooimans, M.A.; Oostra, A.B.; Stumm, M.; Zdzienicka, M.Z.; Joenje, H.; de Winter, J.P. Warsaw breakage syndrome, a cohesinopathy associated with mutations in the XPD helicase family member DDX11/ChlR1. Am. J. Hum. Genet. 2010, 86, 262–266. [Google Scholar] [CrossRef]
  130. Jegadesan, N.K.; Branzei, D. DDX11 loss causes replication stress and pharmacologically exploitable DNA repair defects. Proc. Natl. Acad. Sci. USA 2021, 118, e2024258118. [Google Scholar] [CrossRef]
  131. Calì, F.; Bharti, S.K.; Perna, R.D.; Brosh, R.M., Jr.; Pisani, F.M. Tim/Timeless, a member of the replication fork protection complex, operates with the Warsaw breakage syndrome DNA helicase DDX11 in the same fork recovery pathway. Nucleic Acids Res. 2016, 44, 705–717. [Google Scholar] [CrossRef] [PubMed]
  132. Cortone, G.; Zheng, G.; Pensieri, P.; Chiappetta, V.; Tatè, R.; Malacaria, E.; Pichierri, P.; Yu, H.; Pisani, F.M. Interaction of the Warsaw breakage syndrome DNA helicase DDX11 with the replication fork-protection factor Timeless promotes sister chromatid cohesion. PLoS Genet. 2018, 14, e1007622. [Google Scholar] [CrossRef]
  133. Simon, A.K.; Kummer, S.; Wild, S.; Lezaja, A.; Teloni, F.; Jozwiakowski, S.K.; Altmeyer, M.; Gari, K. The iron–sulfur helicase DDX11 promotes the generation of single-stranded DNA for CHK1 activation. Life Sci. Alliance 2020, 3, e201900547. [Google Scholar] [CrossRef] [PubMed]
  134. Pellicanò, G.; Al Mamun, M.; Jurado-Santiago, D.; Villa-Hernández, S.; Yin, X.; Giannattasio, M.; Lanz, M.C.; Smolka, M.B.; Yeeles, J.; Shirahige, K.; et al. Checkpoint-mediated DNA polymerase ε exonuclease activity curbing counteracts resection-driven fork collapse. Mol. Cell 2021, 81, 2778–2792. [Google Scholar] [CrossRef]
  135. Zhang, Y.; Liu, L.; Wu, X.; An, X.; Stubbe, J.; Huang, M. Investigation of in vivo diferric tyrosyl radical formation in Saccharomyces cerevisiae Rnr2 protein: Requirement of Rnr4 and contribution of Grx3/4 and Dre2 proteins. J. Biol. Chem. 2011, 286, 41499–41509. [Google Scholar] [CrossRef]
  136. Mühlenhoff, U.; Molik, S.; Godoy, J.R.; Uzarska, M.A.; Richter, N.; Seubert, A.; Zhang, Y.; Stubbe, J.; Pierrel, F.; Herrero, E.; et al. Cytosolic monothiol glutaredoxins function in intracellular iron sensing and trafficking via their bound iron-sulfur cluster. Cell Metab. 2010, 12, 373–385. [Google Scholar] [CrossRef]
  137. Netz, D.J.; Genau, H.M.; Weiler, B.D.; Bill, E.; Pierik, A.J.; Lill, R. The conserved protein Dre2 uses essential [2Fe–2S] and [4Fe–4S] clusters for its function in cytosolic iron–sulfur protein assembly. Biochem. J. 2016, 473, 2073–2085. [Google Scholar] [CrossRef] [PubMed]
  138. Zhang, Y.; Li, H.; Zhang, C.; An, X.; Liu, L.; Stubbe, J.; Huang, M. Conserved electron donor complex Dre2–Tah18 is required for ribonucleotide reductase metallocofactor assembly and DNA synthesis. Proc. Natl. Acad. Sci. USA 2014, 111, E1695–E1704. [Google Scholar] [CrossRef]
  139. Li, H.; Stümpfig, M.; Zhang, C.; An, X.; Stubbe, J.; Lill, R.; Huang, M. The diferric-tyrosyl radical cluster of ribonucleotide reductase and cytosolic iron-sulfur clusters have distinct and similar biogenesis requirements. J. Biol. Chem. 2017, 292, 11445–11451. [Google Scholar] [CrossRef]
  140. Furukawa, T.; Naitoh, Y.; Kohno, H.; Tokunaga, R.; Taketani, S. Iron deprivation decreases ribonucleotide reductase activity and DNA synthesis. Life Sci. 1992, 50, 2059–2065. [Google Scholar] [CrossRef]
  141. Sanvisens, N.; Bañó, M.C.; Huang, M.; Puig, S. Regulation of ribonucleotide reductase in response to iron deficiency. Mol. Cell 2011, 44, 759–769. [Google Scholar] [CrossRef] [PubMed]
  142. Tran, P.; Mishra, P.; Williams, L.G.; Moskalenko, R.; Sharma, S.; Nilsson, A.K.; Watt, D.L.; Andersson, P.; Bergh, A.; Pursell, Z.F.; et al. Altered dNTP pools accelerate tumour formation in mice. Nucleic Acids Res. 2024, 52, 12475–12486. [Google Scholar] [CrossRef] [PubMed]
  143. Muñoz, S.; Méndez, J. DNA replication stress: From molecular mechanisms to human disease. Chromosoma 2017, 126, 1–15. [Google Scholar] [CrossRef] [PubMed]
  144. Petronek, M.S.; Spitz, D.R.; Allen, B.G. Iron–sulfur cluster biogenesis as a critical target in cancer. Antioxidants 2021, 10, 1458. [Google Scholar] [CrossRef]
  145. Lee, J.; Roh, J.L. Targeting iron-sulfur clusters in cancer: Opportunities and challenges for ferroptosis-based therapy. Cancers 2023, 15, 2694. [Google Scholar] [CrossRef]
Figure 1. DNA replication and the response to replication stress. The CMG replicative helicase unwinds the parental DNA and allows the recruitment of the Pol α-primase complex through its interaction with Ctf4. The Pol α-primase complex synthesizes RNA primers, which serve as starting points for DNA synthesis. DNA polymerases ε (Pol ε) and δ (Pol δ) catalyze DNA replication on the leading and lagging strands, respectively, by incorporating deoxyribonucleoside triphosphates (dNTPs) produced via the ribonucleotide reductase (RNR)-dependent pathway. When replication fork progression is hindered by obstacles or insufficient dNTP levels, single-stranded DNA (ssDNA) regions are generated and coated by Replication Protein A (RPA). Long RPA-coated ssDNA stretches activate the Mec1/ATR-dependent checkpoint, which delays cell cycle progression, inhibits the firing of late replication origins, upregulates dNTP production and facilitates DNA repair.
Figure 1. DNA replication and the response to replication stress. The CMG replicative helicase unwinds the parental DNA and allows the recruitment of the Pol α-primase complex through its interaction with Ctf4. The Pol α-primase complex synthesizes RNA primers, which serve as starting points for DNA synthesis. DNA polymerases ε (Pol ε) and δ (Pol δ) catalyze DNA replication on the leading and lagging strands, respectively, by incorporating deoxyribonucleoside triphosphates (dNTPs) produced via the ribonucleotide reductase (RNR)-dependent pathway. When replication fork progression is hindered by obstacles or insufficient dNTP levels, single-stranded DNA (ssDNA) regions are generated and coated by Replication Protein A (RPA). Long RPA-coated ssDNA stretches activate the Mec1/ATR-dependent checkpoint, which delays cell cycle progression, inhibits the firing of late replication origins, upregulates dNTP production and facilitates DNA repair.
Cells 14 00442 g001
Figure 2. Biogenesis of cytosolic and nuclear Fe-S cluster proteins in yeast. In the mitochondrion, the mitochondrial Fe-S cluster (ISC) system generates a sulfur-containing compound (X-S) and delivers it to the cytosol, where it becomes a substrate of the cytosolic Fe-S cluster assembly (CIA) system. [4Fe-4S]2+ clusters are assembled on the Nbp35-Cfd1 heterotetrameric scaffold complex, thanks to the electron transfer chain mediated by Tah18 and the Fe-S protein Dre2. Then, Nar1 promotes the transfer of the Fe-S cluster to the CIA targeting complex Cia1-Cia2-Met18. The CIA targeting complex inserts the cluster into the target cytosolic and nuclear apoproteins.
Figure 2. Biogenesis of cytosolic and nuclear Fe-S cluster proteins in yeast. In the mitochondrion, the mitochondrial Fe-S cluster (ISC) system generates a sulfur-containing compound (X-S) and delivers it to the cytosol, where it becomes a substrate of the cytosolic Fe-S cluster assembly (CIA) system. [4Fe-4S]2+ clusters are assembled on the Nbp35-Cfd1 heterotetrameric scaffold complex, thanks to the electron transfer chain mediated by Tah18 and the Fe-S protein Dre2. Then, Nar1 promotes the transfer of the Fe-S cluster to the CIA targeting complex Cia1-Cia2-Met18. The CIA targeting complex inserts the cluster into the target cytosolic and nuclear apoproteins.
Cells 14 00442 g002
Figure 3. Possible roles of Fe-S cluster proteins in yeast replication stress response. Replication stress causes replication fork stalling. The resumption of DNA replication requires the replisome to be maintained on DNA until the stress is removed, the preservation of the replication fork structure, and a sufficient pool of dNTPs to restart and complete DNA replication. As several Fe-S proteins are involved in these processes, the CIA pathway may contribute to overcoming replication stress by regulating some or all of these factors. Specifically, it could help stabilize DNA polymerases on DNA or enhance the activity of nucleases and helicases, such as Dna2, which participate in fork remodeling and the restoration of a functional replication fork. Finally, CIA appears to regulate the nucleotide pool, although the molecular mechanism underlying this regulation remains unclear.
Figure 3. Possible roles of Fe-S cluster proteins in yeast replication stress response. Replication stress causes replication fork stalling. The resumption of DNA replication requires the replisome to be maintained on DNA until the stress is removed, the preservation of the replication fork structure, and a sufficient pool of dNTPs to restart and complete DNA replication. As several Fe-S proteins are involved in these processes, the CIA pathway may contribute to overcoming replication stress by regulating some or all of these factors. Specifically, it could help stabilize DNA polymerases on DNA or enhance the activity of nucleases and helicases, such as Dna2, which participate in fork remodeling and the restoration of a functional replication fork. Finally, CIA appears to regulate the nucleotide pool, although the molecular mechanism underlying this regulation remains unclear.
Cells 14 00442 g003
Figure 4. Regulation of RNR by iron in yeast. Active ribonucleotide reductase (RNR) is formed by R1 and R2 subunits. The Rnr2 R2 subunit carries a Fe3+2-Y∙ cofactor, whose assembly requires the monothiol glutaredoxins Grx3/Grx4 and the electron donor complex Dre2-Tah18. Fe3+2-Y∙ is necessary for RNR catalytic activity and the conversion of ribonucleoside diphosphates (rNDPs) to deoxyribonucleoside diphosphates (dNDPs). Dre2-Tah18 also controls RNR2/RNR4 expression both by counteracting checkpoint activation and the subsequent removal of the transcriptional repressor Crt1, and by reducing the levels of the RNA-binding proteins Cth1/Cth2, which in turn regulate RNR2/RNR4 mRNA stability. Also, the late-CIA complex Cia1-Cia2-Met18 likely participates in the control of dNTP homeostasis, although whether it controls RNR or other steps in dNTP biosynthesis remains to be elucidated.
Figure 4. Regulation of RNR by iron in yeast. Active ribonucleotide reductase (RNR) is formed by R1 and R2 subunits. The Rnr2 R2 subunit carries a Fe3+2-Y∙ cofactor, whose assembly requires the monothiol glutaredoxins Grx3/Grx4 and the electron donor complex Dre2-Tah18. Fe3+2-Y∙ is necessary for RNR catalytic activity and the conversion of ribonucleoside diphosphates (rNDPs) to deoxyribonucleoside diphosphates (dNDPs). Dre2-Tah18 also controls RNR2/RNR4 expression both by counteracting checkpoint activation and the subsequent removal of the transcriptional repressor Crt1, and by reducing the levels of the RNA-binding proteins Cth1/Cth2, which in turn regulate RNR2/RNR4 mRNA stability. Also, the late-CIA complex Cia1-Cia2-Met18 likely participates in the control of dNTP homeostasis, although whether it controls RNR or other steps in dNTP biosynthesis remains to be elucidated.
Cells 14 00442 g004
Table 1. The Fe-S assembly machinery in yeast and human cells.
Table 1. The Fe-S assembly machinery in yeast and human cells.
YeastHumanFunction
ISC system componentsIsuISCUMajor scaffold protein for Fe-S cluster assembly
Nfs1NFS1Cysteine desulfurase that provides sulfur to Isu/ISCU protein
Yah1FDX2Ferredoxin involved in electron transport chain from NADH to sulfur
Arh1FDXRFerredoxin reductase that participates in electron transport chain from NADH to sulfur
Yfh1FrataxinPromotes the delivery of iron to Isu/ISCU protein
Ssq1HSPA9Chaperone that stabilizes scaffold protein and induces cluster release
Jac1HSC20Chaperone that stabilizes scaffold protein and induces cluster release
Isa1ISCA1Synthesizes [4Fe-4S]+ cluster from [2Fe-2S]2+ cluster
Isa2ISCA2Synthesizes [4Fe-4S]+ cluster from [2Fe-2S]2+ cluster
Iba57IBA57Synthesizes [4Fe-4S]+ cluster from [2Fe-2S]2+ cluster
Nfu1NFU1Ensures transfer of Fe-S cluster to mitochondrial targets
Atm1ABCB7ABC transporter that exports X-S from mitochondria
CIA system componentsCfd1CFD1/NUBP2Forms complex with Nb35/NBP35 that coordinates [4Fe-4S]2+ clusters
Nbp35NBP35/NUBP1In complex with Cfd1/CFD1, coordinates [4Fe-4S]2+ clusters
Tah18NDOR1Component of cytosolic electron transport chain
Dre2CIAPINComponent of cytosolic electron transport chain
Grx3-Grx4PICOTPromotes Fe-S cluster assembly on Dre2
Nar1IOP1Adapter between early- and late-acting CIA
Cia1CIA1Inserts Fe-S cluster into target proteins
Cia2CIA2B/FAM96B/MIP18Inserts Fe-S cluster into target proteins
Met18MMS19Inserts Fe-S cluster into target proteins
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Frigerio, C.; Galli, M.; Castelli, S.; Da Prada, A.; Clerici, M. Control of Replication Stress Response by Cytosolic Fe-S Cluster Assembly (CIA) Machinery. Cells 2025, 14, 442. https://doi.org/10.3390/cells14060442

AMA Style

Frigerio C, Galli M, Castelli S, Da Prada A, Clerici M. Control of Replication Stress Response by Cytosolic Fe-S Cluster Assembly (CIA) Machinery. Cells. 2025; 14(6):442. https://doi.org/10.3390/cells14060442

Chicago/Turabian Style

Frigerio, Chiara, Michela Galli, Sara Castelli, Aurora Da Prada, and Michela Clerici. 2025. "Control of Replication Stress Response by Cytosolic Fe-S Cluster Assembly (CIA) Machinery" Cells 14, no. 6: 442. https://doi.org/10.3390/cells14060442

APA Style

Frigerio, C., Galli, M., Castelli, S., Da Prada, A., & Clerici, M. (2025). Control of Replication Stress Response by Cytosolic Fe-S Cluster Assembly (CIA) Machinery. Cells, 14(6), 442. https://doi.org/10.3390/cells14060442

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