Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
J Cell Physiol. Author manuscript; available in PMC 2017 Jan 1.
Published in final edited form as:
PMCID: PMC4994891
NIHMSID: NIHMS810463
PMID: 26040249

Mechanisms and Consequences of Double-strand DNA Break Formation in Chromatin

Abstract

All organisms suffer double-strand breaks (DSBs) in their DNA as a result of exposure to ionizing radiation. DSBs can also form when replication forks encounter DNA lesions or repair intermediates. The processing and repair of DSBs can lead to mutations, loss of heterozygosity, and chromosome rearrangements that result in cell death or cancer. The most common pathway used to repair DSBs in metazoans (non-homologous DNA end joining) is more commonly mutagenic than the alternative pathway (homologous recombination mediated repair). Thus, factors that influence the choice of pathways used DSB repair can affect an individual’s mutation burden and risk of cancer. This review describes radiological, chemical and biological mechanisms that generate DSBs, and discusses the impact of such variables as DSB etiology, cell type, cell cycle, and chromatin structure on the yield, distribution, and processing of DSBs. The final section focuses on nucleosome-specific mechanisms that influence DSB production, and the possible relationship between higher order chromosome coiling and chromosome shattering (chromothripsis).

Keywords: Base excision repair, chromatin, chromothripsis, DNA double-strand break formation, double-strand break repair, homologous recombination, non-homologous end-joining, microhomology-mediated end joining, ionizing radiation, nucleosomes, replication forks, reactive oxygen species

1. MECHANISMS THAT GENERATE DOUBLE-STRAND BREAKS IN DNA

Double-strand breaks (DSBs) in DNA form as a result of exposure to exogenous agents such as radiation and certain chemicals, as well as through endogenous processes, including DNA replication and repair. In addition to these inadvertent occurrences, meiosis I entails the deliberate induction of DSBs, which triggers homologous recombination, thus helping to ensure normal chromosome segregation (reviewed in (de Massy, 2013)). Programmed formation of DSBs also occurs during the development of somatic nuclei in protozoans (reviewed in (Duharcourt et al., 2009)), mating-type switching in yeast (reviewed in (Haber, 2012)), T-cell receptor formation in T-lymphocytes, and immunoglobulin class switching in B-lymphocytes (reviewed in (Soulas-Sprauel et al., 2007)). In this review, we will focus solely on inadvertently produced DSBs. This first section describes the nature and major sources of ionizing radiation (IR), the concept of radiation quality, and mechanisms that link radiation-exposure to the formation of DSBs. We also describe the production of DSBs by radiomimetic chemicals, and the endogenous production of DSBs during DNA replication.

What is ionizing radiation and where does it come from?

IR can be defined as any subatomic particle or electromagnetic wave that possesses enough energy to liberate electrons from atoms. A loss of electrons can disrupt covalent bonds and produce reactive oxygen species (ROS) that can react chemically with, and thereby damage, biological molecules. High-energy protons, electrons, and other charged particles, can dislodge electrons through direct electrostatic (Coulombic) interactions. High-energy but electrically neutral particles, such as neutrons, can interact with atomic nuclei to produce new isotopes. These are commonly unstable (i.e. radioactive), and emit charged α-particles (consisting of two protons and two neutrons) or β-particles (electrons produced by the decay of neutrons into protons, electrons, and neutrinos) as they decay. Photons are electrically neutral ‘wave packets’ that carry the electromagnetic force. Given sufficient energy, they can dislodge electrons via the photoelectric effect (Einstein, 1905).

The particles with the highest energies are found in cosmic rays, which originate mainly from the supernovae of massive stars and from supermassive black holes that reside at the center of the Milky Way and many other galaxies. Virtually all (~99%) of these particles are atomic nuclei stripped of their electron shells and traveling at relativistic speeds. Of these, ~90% are simple protons, derived from hydrogen, while another ~9% are α particles, derived from helium. The remaining ~1% are high-Z, high-energy (HZE) atomic nuclei, derived from heavier elements (Mewaldt, 1994; NASA, 2006). Some of these particles, electrons especially (Baker et al., 2014), are deflected or trapped by the Earth’s magnetic field, forming the Van Allen radiation belts. Particles that penetrate the radiation belts and enter the earth’s atmosphere collide with atomic nuclei in the air. These collisions produce highly energetic protons, antiprotons, and other, much less stable free hadrons, which decay to form electrons, neutrons, protons, α particles, and photons. The resulting cascade of ionized particles and photons, known as an air shower, can be many km wide (Hinton and Hofmann, 2009). Air shower-associated particles with energies that vastly exceed those needed for ionization (in the range of 1 GeV) arrive on earth at a rate of ~10,000 per square meter per second (Pierre Auger Observatory). For comparison, the electrons and protons that make up the solar wind have energies that range between 1.5 and 10 keV. Thus, the sun is responsible for only a small fraction of our cosmic radiation-derived exposure (Kim et al., 1999). Indeed, particles associated with coronal mass ejections (solar flares) may sometimes act as a barrier to lower energy cosmic rays (NASA and Phillips, 2005).

When IR meets biological tissues

As will be detailed in later sections, the type and energy of IR critically affects the kinds of DNA damage created, which in turn may influence the cell’s survival and the means by which the damaged DNA is repaired. IR-associated particles and photons penetrate biological tissues in straight tracks, producing secondary tracks as they interact with atoms in the tissues. Linear energy transfer (LET) is defined as the energy transferred from an IR particle or photon per unit of distance traveled, and is reported as kiloelectron volts per micrometer. As summarized in Table 1, protons, α particles, and high atomic number HZE ions may ionize virtually every atom in their path, giving them high LET coefficients. In contrast, photons and electrons may travel several hundred nanometers between ionization events, giving them much lower LET coefficients. Thus, while protons and α particles will not penetrate as deeply as photons or electrons of equivalent energies, they are far more damaging per unit volume of tissue penetrated.

TABLE 1

Summary of ionizing radiation types, “quality” (i.e. low or high LET), and damage hallmarks.

Ionizing
radiation
type
Common
sources
Linear
Energy
Transfer
(LET)*
%
Annual
Exposure
**
Particle
track
Damage
clustering
Damage
Hallmarks
PhotonsGamma
rays &
X-Rays
Low~35%High
penetration,
crooked
track
≤10
lesions per
cluster
Highly penetrating
but sparse
damage;
Random
distribution of
DSBs
β
particles
ProtonsCosmic
rays &
radon
emissions
High~65%Low
penetration,
straight
track
≤25
lesions per
cluster
Complex clustered
lesions
surrounding DSB;
“Dirty breaks”;
Non-random DSBs
(average break
size 2kbp);
Lagging repair
α
particles
HZE ions
*While photons cannot directly be described by LET, the secondary electrons (β particles) created by them will have low LET.
Theoretically determined by Monte Carlo algorithm (Semenenko and Stewart, 2004)

High LET radiation is a particular risk to astronauts as well as cancer patients treated with heavy ion radiation. Astronauts orbiting the earth are protected from cosmic rays, to a degree, by the Van Allen Radiation belts, whereas those traveling to the moon or beyond are at greater risk. There is considerable uncertainty about the actual magnitude of the increased radiation risk (NASA, 1998), due mainly to sparse information on the biological responses to high LET radiation. The cosmic ray flux, measured on Earth, varies considerably with latitude and longitude, mainly because the Earth’s magnetic field diminishes near the north and south poles. As a results, cosmic ray-derived IR may thus account for a large fraction of the total radiation exposure for airline flight crews that routinely fly polar routes. However, radon accounts for a much larger fraction (~42%) of the general population’s annual exposure to high LET radiation. Radon forms as an intermediate in the decay of thorium and uranium into lead, and is the only gas that consists entirely of radioactive isotopes. Other radioactive isotopes present in food, soil and building materials, account for another ~26% of one’s average annual IR exposure.

The largest source of low LET IR comes in the form of high-energy photons, although β particles produced by radioactive decay also contribute to low LET IR in the environment. A photon’s energy is inversely proportional to its wavelength, and photons that carry visible light (wavelengths of ~380 to ~750 nm) are not energetic enough to ionize atoms. Longer wavelength photons, such as those used in microwaves and transmission of radio and television signals, carry even less energy. The shorter wavelength photons associated with the other side of the visible light spectrum include ultraviolet rays, X-rays, and gamma rays, and carry much higher energies. Photons with wavelengths shorter than ~100 nm are energetic enough to ionize hydrogen atoms and break chemical bonds. The energies associated with airport security, medical X-rays and CT scans can be 1,000 to 10,000 times higher than this. Collectively, these devices are estimated to account for ~20% of the average person’s annual IR exposure.

Direct induction of double-strand breaks by IR

By the late 1930s, experimental geneticists had established that IR can break chromosomes (Sax, 1938). With the advent of methods to isolate very long DNA, it became possible to demonstrate, using velocity sedimentation of DNA through neutral and alkaline sucrose gradients, that IR breaks the DNA within chromosomes. These breaks can occur either directly or indirectly. The most direct path entails a collision between a high-energy particle or photon and a strand of DNA, breaking the phosphodiester backbone. More commonly, IR splits water molecules near DNA, creating hydrogen and hydroxyl free radicals ((Yamaguchi et al., 2005) and reviewed in (Barnard et al., 2013; Ward, 1988)). The short-lived, but highly reactive hydroxyl radicals may react with nearby DNA, producing single-strand DNA breaks (SSBs). Closely opposed SSBs, created by either route, may spontaneously convert into a double-strand break (DSB). In the simplest version of the SSB to DSB conversion hypothesis, each SSB forms independently. The number of SSBs that form each day in mammalian cells has been estimated at ~55,000 (Tice and Setlow, 1985), roughly one SSB per ~105 bp, or one SSB per ~104 helical turns. The probability of two opposing SSBs forming within one or two DNA helical turns of one another would thus be on the order of (10−4 × 10−4) = 10−8. This frequency could account for the estimated 10–50 DSBs that form every day in every nucleated human cell ((Vilenchik and Knudson, 2003); and reviewed in (Mehta and Haber, 2014)), but only if SSBs persist for a large fraction of the cell cycle. As described below, some do but many do not. The studies that called into question the simple SSB to DSB conversion hypothesis tested the prediction that a 10-fold increase in rates of SSB formation, (e.g. from 10−4 to 10−3 per one or two helical turns) would be expected to increase the rate of DSB formation by ~100-fold (from 10−8 to (10−3 × 10−3) = 10−6). The predicted quadratic relationship proved true for peroxide-generated DSBs but not for radiation-induced DSBs, which increased linearly with radiation doses up to several hundred Gray (Dahm-Daphi et al., 2000). The simplest interpretation of these results was that each peroxide-generated SSB reflects a single independent event, but that a collision with a single IR particle generates multiple damages, including closely opposed SSBs. These closely opposed lesions, initially dubbed “locally multiply damaged sites”, or LMDSs (Ward, 1988), are now commonly referred to as “clustered damages.” They are defined as two or more DNA lesions, created by a single track of radiation, that reside within one or two helical turns of DNA. Low or high LET radiation doses as low as 1 Gray (100 rad) can produce clustered lesions (Goodhead et al., 1993; Rydberg, 1996; Sutherland et al., 2000a), and Monte Carlo-based modeling of radiation tracks suggests that low and high LET radiation can generate, respectively, up to 10 and 25 lesions per damage cluster (Semenenko and Stewart, 2004). These may hinder the DNA processing events associated with DSB repair (see “A short primer on double-strand break repair,” below). As well, DNA ligase cannot reseal a SSB unless the break is “clean,” that is consisting of one 3’ hydroxyl end and one 5’ phosphate end. “Dirty” single and double strand DNA breaks created by high or low LET IR may contain virtually unprocessable, unligatable ends (Weinfeld and Soderlind, 1991). Tyrosyl-DNA phosphodiesterase, for example, can require several hours to remove phosphoglycolate moieties from 3’ DNA ends (Zhou et al., 2009).

Indirect generation of DSBs during attempted base excision repair (BER) of clustered lesions

Although the above-described direct mechanisms account for many of the DSBs that form in cells, a significant fraction is generated by indirect mechanisms. The reactive oxygen species (ROS) produced by IR-mediated radiolysis of water generate not only SSBs, but also a wide array of other DNA damages, including oxidized bases and sites of base loss (De Bont and van Larebeke, 2004). Most such lesions are subject to base excision repair (BER), which entails production of gapped repair intermediates (for reviews, see (David et al., 2007; Duclos et al., 2012; Hegde et al., 2008; Krokan and Bjoras, 2013; Robertson et al., 2009; Wallace, 2014)). Normally, these 1–2 nucleotide DNA gaps are filled by DNA polymerase β and sealed by DNA ligase IIIα. However, the near-simultaneous attempted BER of clustered oxidative lesions can produce SSBs in both DNA strands, which may convert to a DSB, as illustrated in Figure 1. BER-dependent formation of DSBs has been observed in multiple in vitro studies (for reviews, see (Eccles et al., 2011; Sage and Harrison, 2011)) and now forms the basis for assays to measure the abundance of non-strand break clustered damages (Sutherland et al., 2000b; Sutherland et al., 2003). Specifically, one treats oxidatively damaged DNA with a mixture of recombinant, prokaryote-derived DNA glycosylases and endonucleases, fractionates the resulting DNA fragments by gel electrophoresis, and then calculates the break frequency from the number average size of the DNA fragments. There is compelling evidence that BER-driven DSB formation also occurs in vivo, in both prokaryotes and eukaryotes (Blaisdell et al., 2001; Blaisdell and Wallace, 2001; Yang et al., 2006; Yang et al., 2004). Specifically, mutation or knockdown of DNA glycosylases reduced the frequency of DSB formation in cells that had been exposed to IR. Conversely, overexpression of those same glycosylases sensitized cells to radiation and increased DSBs. The section below, “Impact of chromatin on the formation and persistence of DSBs”, describes mechanisms by which nucleosomes suppress (but do not fully eliminate) BER-dependent production of DSBs in eukaryotes.

An external file that holds a picture, illustration, etc.
Object name is nihms810463f1.jpg
DSB formation via attempted base excision repair of closely opposed oxidative lesions

Reactive oxygen species (ROS) can produce oxidized DNA bases, sites of base loss and DNA nicks. The base excision repair (BER) of single lesions (left) begins with the excision of oxidized bases (Ox) by DNA glycosylases, or with the excision of apurinic/apyrimidinic (AP) sites by apurinic endonuclease. In either case, apurinic endonuclease goes on to generate a single base gap that is filled by DNA polymerase β and sealed by DNA ligase IIIα, in complex with XRCC1. Because a single IR particle can produce multiple clustered ROS, IR often generates a cluster of oxidative lesions in DNA. Near-simultaneous BER of closely opposed lesions in such a cluster can generate closely spaced nicked or gapped repair intermediates in opposing DNA strands. These may spontaneously convert to DSBs before BER is complete (right).

DNA damaging ROS are produced not only by IR-mediated radiolysis of water but also during normal oxidative metabolism. Most of these endogenously-generated ROS are neutralized, but enough escape to generate roughly 10,000 oxidative base lesions and another 10,000 base loss events per cell per day ((Fraga et al., 1990; Lindahl and Nyberg, 1972) and reviewed in (Friedberg, 2006)). Given this, one might ask why the frequency of DSB formation is no higher than observed. The answer seems to be that, unlike the multiple, highly localized ROSs produced by a single IR particle or photon, each endogenously-generated ROS is produced independently and, therefore, rarely form clustered lesions.

DNA replication-associated DSBs

Most of the DSBs that can be attributed to endogenous processes are produced during DNA replication (Syeda et al., 2014). As depicted on the left in Figure 2, a replicative polymerase that encounters a SSB in the template strand may stall, leading to a collapse of the replicative fork and subsequent DSB formation (Pfeiffer et al., 2000). The SSB itself may be present as a DNA repair intermediate. DSBs can also occur as a result of replication fork stalling due unusual DNA secondary structures, bulky lesions, polymerase blocking oxidative lesions, abasic sites, chemical or IR-generated inter-strand crosslinks, or as a result of collisions with transcription complexes and certain DNA binding proteins (Dextraze et al., 2010; Mirkin and Mirkin, 2007; Prado and Aguilera, 2005). As depicted on the right in Figure 2, stalled replication forks may regress and partially displace newly synthesized DNAs from their template strands, enabling the 3’ end of the leading strand to anneal to the 5’ end of the lagging strand. The resulting “chicken foot” is structurally identical to a Holliday junction, which can be cleaved by Holliday junction resolvases or structure-directed nucleases, forming a DSB in the process (reviewed in (Mehta and Haber, 2014)). Importantly, DSBs generated at collapsed replication forks may sometimes be repaired by the highly mutagenic microhomology-mediated end-joining pathway (see section “A short primer on double-strand break repair”) rather than by error-free recombination mediated repair (Truong et al., 2013).

An external file that holds a picture, illustration, etc.
Object name is nihms810463f2.jpg
DSB formation via replication fork collapse

(Left) In the simplest scenario, leading strand synthesis is halted by a single-strand break (SSB), created via exogenous damaging agents or as a repair intermediate (see Figure 1). Collapse of the replication fork converts the SSB into a one-sided DSB. The attempted repair of one-sided DSBs by NHEJ would potentially lead to chromosomal rearrangements or translocations. More commonly, one-sided DSBs may initiate break-induced DNA replication (reviewed in (Malkova and Ira, 2013)). (Right) Several kinds of DNA lesions, including thymine dimers, certain oxidized bases (e.g. thymine glycol), abasic sites, and inter-strand crosslinks (such as those caused by cisplatin) can cause replication forks to stall. If the stalled replication fork regresses, it will partially displace newly synthesized leading and lagging strands, allowing them to anneal, as depicted. The newly synthesized lagging strand may then serve as a template to further extend the leading strand, producing the “chicken foot” intermediate shown. This may resolve if the replication block is removed, allowing replication to restart. Alternatively, because the chicken foot is structurally analogous to a Holliday Junction, it may be cleaved by resolvases, producing, once again, a one-sided DSB.

Transcription and R-loop-related formation of DSBs

Numerous studies have reported higher than average mutation rates in heavily transcribed regions of the genome. Some of these mutations may be due to elevated rates of DSB formation in these same regions. Variable rates of DSB formation across the genome are influenced by chromatin structure, as discussed in Section 3, below. However, some DSBs in heavily transcribed regions appear to result from transcription per se. DNA unwinding during the formation of replication and transcription bubbles generates superhelical stress in flanking DNA, which is relieved by topoisomerases. Once formed, transcription bubbles, and the ~8 bp RNA-DNA hybrids that form within them, remain relatively constant in size during transcription, because DNA unwinding ahead of the elongating RNA polymerase is matched by the progressive displacement of nascent RNA and DNA rewinding behind the RNA polymerase. However, transcription elongation may occasionally outpace the action of topoisomerases, resulting in a transient accumulation of positive superhelical density ahead of the RNA polymerase, and a corresponding accumulation of negative superhelical density in the trailing region (Liu and Wang, 1987). Negative superhelical density may delay reannealing of DNA, and enable a portion of the nascent RNA to reanneal to its DNA complement, after it exits the RNA polymerase. The result is an R-loop. At least in yeast, R-loop formation also requires Rad51p and Rad52p, which normally act in homologous recombination, and is suppressed by the DNA helicase and Rad51p antagonist, Srs2p (Wahba et al., 2013). One might expect the packaging and processing of nascent RNAs to reduce the probability of R-loop formation, and RNA-DNA helicases and RNase H enzymes to restrict the lifetime of R-loops that do form. Consistent with this prediction, defects in certain RNA processing enzymes produce hyper-recombination phenotypes can be suppressed by overexpression of RNase H (reviewed in (Hamperl and Cimprich, 2014)). Exactly how R-loops are transformed into DSBs is unclear, but candidates include the DNA and RNA editing enzymes in the apolipoprotein B mRNA editing enzyme, catalytic polypeptide-like (APOBEC) family (Hamperl and Cimprich, 2014). The prototype enzyme in this family is activation-induced cytidine deaminase (AID), which converts cytidine to uracil during class switching in B cells (Stavnezer, 2011). The resulting dU-dG mismatches are substrates for uracil-DNA glycosylase (UDG), which excises uracil moieties, leaving an abasic site. The abasic site could subsequently halt progression of a replication fork, leading in some cases to fork collapse and rescue via homologous recombination. Alternatively, the abasic site might be processed further via BER, with formation of a single-strand gapped intermediate that might then be converted to a DSB during replication. The postulated involvement of DNA replication in R-loop mediated DSB formation is consistent with the observation that the hyper-recombination phenotype seen in certain RNA processing mutants is confined to genes transcribed during S phase, and is not evident among genes transcribed in G2 phase (Wellinger et al., 2006). Various genome wide genetic and proteomic screens have further cemented the links between defective RNA processing, R-loop formation, and genome instability. For example, certain RNA processing factors are among those proteins targeted by the DNA damage response kinases, ataxia telangiectasia mutated (ATM) and ataxia telangiectasia and Rad3 related (ATR) (Matsuoka et al., 2007) (Smolka et al., 2007).

Induction of DSBs by radiomimetic compounds

Exogenous chemicals that damage DNA are referred to as clastogens. These include anti-cancer drugs, such as the DNA-alkylating agents methyl methanosulfonate and temozolomide, and the DNA cross-linking reagents cisplatin and mitomycin C. Clastogens can induce DSBs indirectly, via the above-described cellular processes. Chemical agents that induce strand breaks directly are commonly referred to as “radiomimetic” compounds because of their ability to induces DSBs. Bleomycin, for example, is a radiomimetic compound that the World Health Organization has included in its 2013 list of essential medicines for treatment of high priority diseases. “Activated bleomycin” contains a reduced transition metal (Fe(II) or Cu(I)), oxygen and a one-electron reductant (often hydrogen peroxide). Decomposition of activated bleomycin in the presence of DNA can generate strand breaks and create additional hydroxyl radicals. Although bleomycin carries just one reactive moiety, it somehow produces about one DSB for every 10 single-strand breaks it generates (Chen and Stubbe, 2005; Povirk, 1996).

Even more efficient than bleomycin at generating DSBs are the bicyclic enediynes, molecules that contain a C-C double bond flanked by C-C triple bonds. These molecules can cyclize to form a 1,4-dehydrobenzene intermediate. Neocarzinostatin, which is produced by Streptomyces carzinostaticus, is a small protein with a tightly bound chromophore. Cyclization of the chromophore forms a highly reactive para-benzyne diradical that can abstract hydrogens from the C-1’, C-4’ and C-5’ positions of deoxyribose moieties in both DNA strands. Subsequent reactions between these carbon atoms and molecular oxygen can produce DSBs. Several enediyne compounds are in clinical trials for possible use in cancer therapies. However, until it becomes possible to deliver these compounds to specific cellular targets, it is unlikely that their use will supplant radiation therapy.

Topoisomerases play critical roles in DNA metabolism, relieving local super-helical stress associated with formation of transcription and replication bubbles, and with the assembly and disassembly of nucleosomes. Class I topoisomerases generate transient SSBs in DNA, while Class II topoisomerases generate transient DSBs, which enables them to carry out decatenation reactions. Camptothecins are topoisomerase I inhibitors, while etoposides are topoisomerase II inhibitors (Hande, 1998; Koster et al., 2007). Both act by stabilizing topoisomerase-DNA intermediates, preventing re-ligation. The topoisomerases can thus be trapped through covalent linkages with DNA ends, creating protein-associated SSBs or DSBs. The protein-DNA moieties must be removed by specific endonucleases before repair can ensue. Thus, repair of such damages necessarily entails deletion of short segments of DNA (Takahashi et al., 2011).

Can our knowledge of mechanisms of DSB formation help guide public policy on radiation exposure?

There is widespread agreement among scientists and policy experts about the importance of limiting exposure to high doses of IR. There is much less agreement on the risks associated with very low dose IR (e.g. (Doss et al., 2014)). The adoption of assays based on counting radiation-induced DSB repair foci, visualized by fluorescent tagging of antibodies to phosphorylated histone H2AX (γ-H2AX), tumor suppressor P53 binding protein 1 (53BP1), or auto-phosphorylated ATM, has made it possible to investigate the occurrence and rates of repair of DNA damage resulting from radiation doses far lower than were commonly used in early studies. With the development of biomarkers for radiation exposure it should be possible to gather more precise epidemiological information (Pernot et al., 2012). However, the number of potentially confounding variables has increased along with the precision and quality of exposure and repair assays. For example, cardiovascular disease may increase mortality following radiation exposure (Little et al., 2012). It is also likely that cancer risks associated with lose dose IR will vary with genetically-determined differences in DNA repair pathways and immune surveillance. Such variables should be considered in the design and evaluation of studies of low dose IR risks. Unfortunately, in doing so, one may lose the very statistical power needed to assess risk. Because of this conundrum, it is unlikely that debates on the risks of low dose IR will be settled anytime soon.

2. A SHORT PRIMER ON DOUBLE-STRAND BREAK REPAIR

Although this review focuses on mechanisms of DSB formation, the impact of DSBs, and the manner in which they are repaired, varies with their etiology, for example whether DSBs result from high or low LET radiation. To facilitate this discussion, this section briefly summarizes the damage response pathways that process DSBs. There are several excellent, recent reviews of DSB repair for readers looking for more detailed expositions (e.g. (Asaithamby and Chen, 2011; Lieber, 2010; Mehta and Haber, 2014; Polo and Jackson, 2011; San Filippo et al., 2008; Thompson, 2012)).

Non-homologous end-joining (NHEJ)

NHEJ is the only DSB pathway that normally operates in G0 and G1 phases of the cell cycle. NHEJ is initiated by the binding of Ku70/80 heterodimers to DNA ends created by a DSB event. Ku heterodimers exist as preformed rings that encircle 3–4 bp of DNA, fitting along the contours of the major and minor grooves (Walker et al., 2001). Once bound to DNA, Ku recruits the X-ray repair cross-complementing protein 4 (XRCC4) and the Cernunnos-X4-like factor (Cer-XLF), which help bridge the two DNA ends. If the DNA ends are complementary and undamaged, they can be joined directly by the XRCC4-associated DNA ligase IV (Reynolds et al., 2012). If the DNA ends are damaged or not fully complementary, ligation cannot occur without further end processing. In this case, Ku recruits DNA protein kinase (DNA-PK). The binding of DNA-PK activates its catalytic subunit (DNA-PKcs), which leads to the recruitment and phosphorylation of the endonuclease Artemis. Provided that one or two complementarity nucleotides in the two DNAs are able to anneal with one another, Artemis will remove excess single-strand DNA. This generates a substrate that can be ligated by DNA ligase IV, acting in concert with XRCC4 and Cer-XLF.

Homology-directed recombination-mediated repair (HRR)

A large fraction of the DSBs that form during the S or G2 phase of the cell cycle are repaired by HRR, which begins with resection of 5’ DNA ends that flank a DSB, as illustrated in Figure 3. This resection is catalyzed primarily by MRN, a trimeric complex containing MRE11, RAD50, and Xrs2/NBS1. MRE11 possesses both single-strand endonuclease and 3’ to 5’ exonuclease activities that can process ‘clean’ DNA ends, such as those created by endogenous expression of restriction endonucleases. However, the processing of DNA ends that form hairpins or are attached to proteins requires additional help from the endonuclease Sae2/Ctp1/CtIP. Acting together, MRE11 and CtIP remove 5’ strands in 50 to 100 NT increments (reviewed in (Mimitou and Symington, 2009)). The single-strand endonuclease DNA2 or the FLAP endonuclease EXO1, acting in concert with the RecQ family helicase Sgs1/BLM, catalyze further resection. The 3’ ssDNA tails generated by these resection events are bound first by the ssDNA binding protein RPA. RPA is subsequently replaced by the RAD51 recombinase, which catalyzes the search for complementary DNA, usually in the form of a sister chromatid. As the RAD51–3’ single-strand DNA filament from one side of the DNA break invades and anneals to complementary sequences in its sister chromatid, it displaces the opposing strand in the sister chromatin, forming a D-loop. The invading 3’ single-strand DNA then primes new DNA synthesis. This extends the D-loop, exposing bases that can anneal to the 3’ single-strand DNA from other side of the DNA break. This second annealing event, when it occurs, primes leading strand synthesis of DNA on the opposing strand. Ultimately, these cross-over structures (“double Holliday junctions”) are resolved, leaving widely separated, single-strand DNA nicks that are sealed by DNA ligase. In some instances, the first invading strand primes new DNA synthesis, but the second annealing event fails to occur. Instead, the original invading strand is displaced and, having been lengthened by new DNA synthesis, can now anneal to 3’ single-strand DNA from other side of the DNA break (Figure 3E). This synthesis-dependent strand annealing (SDSA) provides a template for repair synthesis. HRR is commonly viewed as an error-free repair pathway, and indeed one is less likely to observe the highly localized short deletions and base changes that often accompany NHEJ. On the other hand, the fidelity of DNA repair synthesis is not as high as one sees in normal genomic replication; thus one may find more broadly distributed base substitutions in the wake of HRR (Malkova and Haber, 2012).

An external file that holds a picture, illustration, etc.
Object name is nihms810463f3.jpg
Homology-directed recombination-mediated repair (HRR) via synthesis-dependent strand annealing (SDSA)

The Figure depicts key factors and the sequence of steps in SDSA, a common homology-mediated DSB repair sub-pathway. For clarity and simplicity, other HR-initiated pathways, accessory proteins and replication machinery are not shown, but are described in (Mehta and Haber, 2014; San Filippo et al., 2008) and other HRR-related reviews. Step (A) depicts the initiation of HRR by the MRN complex which, together with the endonuclease Sae2/Ctp1/CtIP resects 5’ DNA ends. This creates 3’ single strand DNA (ssDNA) tails that are bound by the ssDNA binding protein RPA (step (B)). RPA is then replaced by RAD51, which catalyzes a homology search; annealing of the Rad51-ssDNA filament to its homolog displaces the homolog’s normal complement, creating a “D-loop” (step (C)). Using the sister chromatid as a template, the invading strand is able to prime DNA synthesis, thereby extending the D-loop (step (D)). Further extension of the D-loop may enable it to anneal with the second 3’-ssDNA tail, which could then prime DNA synthesis in the opposite direction (not shown). Alternatively, the original, newly-extended invading strand may be displaced, allowing it to anneal with the second 3’-ssDNA tail (step (E)). DNA synthesis primed by the second 3’-ssDNA tail, followed by ligation, would then complete the repair (step (F)).

Alternative (microhomology-mediated) end-joining (alt-EJ/MMEJ)

Microhomology-mediated end-joining is linguistically related to NHEJ but shares key mechanistic features with HRR ((Truong et al., 2013) and references therein). MMEJ appears to operate mainly during S-phase of the cell cycle, in a Ku- and DNA-PK-independent fashion. As with HRR, MMEJ begins with the resection of 5’ DNA ends at the DSB, leaving behind 3’ single-strand DNA. Although MMJE and HRR employ the same end processing enzymes, subsequent steps diverge. MMEJ occurs when end-resection exposes micro-homologies of 5–25 bp that enable DNA single strands to anneal. This creates a substrate that, following removal of non-annealed DNA ends, is competent for gap filling and ligation (by DNA ligase IV in yeast and DNA ligase III in mammalian cells), as also occurs during single strand annealing repair (see (Frankenberg-Schwager et al., 2009)).

Regulation of pathway choice during DSB repair

The pathway that cells use to process DSBs depends on multiple variables, including the cell cycle stage during which the DSB is generated, and whether the DSB itself is ‘clean’ or ‘dirty.’ DSB formation during late S-phase or G2 phase triggers a DNA damage checkpoint that prevents cells from progressing through mitosis until DNA is repaired. Likewise, DSB formation during G1 phase triggers a checkpoint that prevents cells from entering into S phase. Interestingly, the G1/S phase checkpoint is absent in mouse embryonic stem (ES) cells and, as a result, ES cells that suffer DSBs during G1 phase commonly undergo apoptosis; DSBs that form in ES cells during S or G2 phases are repaired by HRR. This may explain why ES cells are hypersensitive to IR, and also why surviving cells accumulate far fewer mutations than do isogenic embryonic fibroblasts (Tichy et al., 2010). In addition to checkpoint control-related mechanisms, the choice of HRR over NHEJ in S and G2 phases is influenced by the cell cycle-dependent, activating phosphorylation of the resection factor Sae2 (Huertas et al., 2008). Presumably, the absence of active Sae2 in G0 or G1 phase suppresses attempted HRR at a stage where cells lack the homologous DNA that HRR requires. It is not clear if there are analogous mechanisms that suppress NHEJ during S and G2 phase, beyond the fact that end-resection at DSBs inhibits the binding of Ku proteins that initiate NHEJ. Interestingly, the checkpoint kinase ATR, when activated, modulates the relative activity of EXO1 and CtIP: ATR-mediated phosphorylation of EXO1 renders it vulnerable to proteolysis, whereas ATR-mediated phosphorylation of CtIP is required for its chromatin binding and resection activity (Peterson et al., 2013). These phosphorylation thus events may influence the choice between MMEJ and HRR. Replicative age may also influence pathway choice. In aging mice, the frequency of MMEJ appears to increase while that of NHEJ declines (Vaidya et al., 2014).

Impact of DSB etiology on DSB repair

Whether DSBs are ‘clean’ or ‘dirty’ can influence both the pathway used to repair a DSB, and the overall rate of repair. As noted earlier, DNA that flanks DSBs produced by high-energy protons and α particles may contain numerous additional damages that slow or inhibit the processing of DNA ends for NHEJ. (Pinto et al., 2005), for example, reported that some DSBs generated by exposure to relatively massive α particles had half-lives of one hour or less, whereas others had half-lives of 12–16 hours; a small fraction of the DSBs persisted so long as to appear unrepairable. (Staaf et al., 2012) reported that gamma (photon) radiation generates as many or more γ-H2AX-marked DSB repair foci than do α particles. However, the rates of repair were faster for gamma radiation-induced DSBs than for α particle-induced DSBs. Taken together, these studies suggest that lesions induced by high LET radiation, even if no greater in total extent, are generally more difficult to repair, making them all the more damaging. For DSBs that form during late S or G2 phase, this may shift the balance toward homologous recombination-mediated repair (or to MMEJ), owing to the capacity of the CtIP endonuclease to bypass complex DNA structures. Because HR requires cells to enter into S or G2 phase, one might predict that complex DSBs that form in G0 or G1 can be repaired only if cells are able to bypass the G1/S phase checkpoint. It appears, however, that CtIP and other end resection factors can be recruited to complex DSBs in G1 phase (Averbeck et al., 2014; Yajima et al., 2013) where, presumably, they initiate MMEJ-mediated repair of complex damages. Interestingly, CtIP must be phosphorylated to act in HRR, but apparently not when acting in MMEJ (Yun and Hiom, 2009).

DSB repair mechanisms and the optimization of radiotherapy for cancer patients

As described above, DSBs induced by high LET radiation are difficult to repair, and may thus be more frequently lethal than DSBs generated by equivalent doses of low LET radiation. However, these predicted differences in cell lethality do not seem to have translated to significantly better outcomes for radiotherapy patients (Terasawa et al., 2009). There may be several reasons for this, including differences in IR dose fractionation and delivery. For example, the high LET coefficients of protons has meant using proton beam energies of 70–250 MeV to ensure sufficient penetration, as compared to photon beam energies of 6–15 MeV for equivalent tumors (e.g. (Sung et al., 2012)). On a more optimistic note, advances in our understanding of DSB repair mechanisms are beginning to pay dividends in the form of adjuvant therapies for treatment of certain cancers. For example, cancer cells containing BRCA 1 or BRCA 2 mutations are generally unable to repair DSBs by HRR, making them heavily reliant on NHEJ and MMEJ. Such cells exhibit elevated vulnerability to IR when treated with inhibitors of poly-ADP ribose polymerase 1 (PARP1), which is required for MMEJ. Accordingly, some PARP1 inhibitors have already been approved for treatment of certain cancers, and others are in phase III clinical trials (Sonnenblick et al., 2015).

3. IMPACT OF CHROMATIN STRUCTURE ON DSB FORMATION

Nucleosome structure and changes in the chromatin landscape as a function of transcription and cell cycle progression

Most of the DNA in nuclei is packaged in chromatin, which consists of regularly spaced nucleosomes, punctuated by nucleosome-free DNA segments that associate with factors engaged in transcription, DNA repair, DNA replication, and telomere-related activities. Each nucleosome consists of ~147 bp DNA, wrapped in a left-handed toroidal helix around a protein core, containing two copies each of histones H2A, H2B, H3 and H4 (Luger et al., 1997). Individual nucleosomes are separated from one another by segments of linker DNA that, in humans, average ~50 bp in length, but can vary from ~20 bp (e.g. in S. cerevisiae) to ~110 bp (e.g. in sea urchin sperm). Most nucleosomes assemble during replication, in a stepwise fashion that is coordinated with movement of the DNA replication fork and facilitated by specific histone chaperones. In the first step, DNA associates with a histone H3–H4 tetramer. This is followed by the addition of histone H2A–H2B dimers to either side of the central H3–H4 tetramer. Prior to and following assembly of nucleosomes, histones are subject to numerous (and in most cases reversible) secondary modifications that influence interactions between nucleosomes and various regulatory and structural factors. In addition to functional heterogeneity imparted through histone modifications, cells contain chaperones that catalyze the exchange of histone primary sequence variants in a replication-independent fashion. Chromatin associated with transcriptionally active regions is generally in a relatively “open” or extended configuration, commonly referred to as a 10 nm filament (or fiber) (Finch and Klug, 1976; Olins and Olins, 1974; Oudet et al., 1975; Woodcock et al., 1976). 10 nm chromatin fibers can coil to produce more compact, 30 nm filaments that, generally, are transcriptionally inactive (Finch and Klug, 1976; Olins and Olins, 1979; Pooley et al., 1974; Ris and Kubai, 1970; Worcel and Benyajati, 1977). In advance of mitosis or meiosis, the 30 nm fibers assemble into radially arrayed 50–100 kbp loops, each anchored to a central axis that forms the long axis of meiotic and mitotic chromosomes. Clearly, chromatin is a dynamic entity, and one might expect its impact on DSB formation to vary across the genome (as a function of transcription, for example) and at different points in the cell cycle. We discuss these variables below.

Nuclear factors and conformational variables in DNA that influence the susceptibility of DNA to free radical-mediated damage

The direct production of DNA strand-breaks is commonly driven by reactions between hydroxyl radicals (OH·) and hydrogen atoms in deoxyribose (Balasubramanian et al., 1998). As summarized in Table 2, those atoms are less accessible in DNA with unusually narrow minor grooves (Isabelle et al., 1995). Variations in groove width are sometimes sequence-directed, which may, for example, explain why 5’-AATT-3’ sequences are less sensitive than average to radiation-induced strand-breaks (Franchet-Beuzit et al., 1993; Sy et al., 1997). In other instances, variations in groove geometry are induced by DNA circularization, supercoiling, or bending, as when DNA winds about the histone octamer. This might explain why plasmid DNA is more radiosensitive in a relaxed configuration than when supercoiled (Swenberg and Speicher, 1995). As well, small molecules, such as the polyamine spermine, can moderate reactions between ROS and DNA (Chiu and Oleinick, 1998; Spotheim-Maurizot et al., 1995; Warters et al., 1999). Spermine may exert this protective effect by displacing water molecules from DNA, thereby reducing the local yield of hydroxyl radicals formed by collisions between IR and water (van Dam et al., 2002); spermine may also react with, and thereby neutralize, local ROS. In a similar fashion, histones and other chromatin associated proteins protect DNA by physically masking reactive moieties, and by scavenging or reacting with ROS. Evidence of this includes studies in which the frequency of DSB formation increased after chromatin-associated proteins were stripped from DNA, using high salt (reviewed in (Lavelle and Foray, 2014)). Although results from such studies vary with methodology, cell type, and radiation dose, they suggest that non-histone proteins and the linker histone H1 provide a two to five-fold protection from DSBs, while the core histones provide the most protection: the frequency of DSBs increased on the order of ~20–40-fold with the removal of histone H2A–H2B dimers, and up to 70-fold with the removal of all four core histones.

TABLE 2

Multiple levels of protection against DSBs.

Compaction
level
Structural factors that suppress
reactions between DNA and
reactive oxygen species (ROS).
Structural factors that suppress
formation of DSB during
replication, transcription, or
attempted base excision repair
(BER).
DNA alone
(and DNA-
associated
small
molecules)
Narrow minor grooves (created by
either sequence context or DNA
bending) limit access by ROS.

Spermine and similar small molecules
may partially displace water from
DNA, reducing the number of
molecules that can be transformed by
IR into ROS. They also provide
reactive moieties that may help
quench local ROS.
Local, sequence-directed bending of
DNA can influence the relative rates
of lesion discovery and repair at sites
containing multiple clustered
damages. In principle, this might
reduce the probability of DSB
formation from near-simultaneous
generation of nicked or gapped repair
intermediates in opposing DNA
strands.
NucleosomeMinor groove narrowing imposed by
wrapping of DNA about the histone
octamer, and histones themselves
partially limit access to ROS reactive
moieties in DNA; histones also
provide reactive moieties that may
help quench local ROS.
As depicted in Figure 4, nucleosomes
substantially restrict the fraction of
clustered oxidative lesions that can be
transformed into DSBs during
attempted BER.
Hetero-
chromatic
regions &
higher order
chromatin
structures*
Linker histones and other chromatin-
associated proteins limit nucleosome
mobility, increasing the total fraction
of DNA protected by histones. Higher
order coiling (particularly the ‘zig-
zag’ or crossed-linker configurations)
may reduce access of ROS to linker
DNA.
Linker histones and other chromatin-
associated proteins that limit
nucleosome mobility may partially
suppress BER-dependent formation
of DSBs while higher order coiling
may abolish BER altogether.
Transcription-associated formation of
DSBs would be similarly affected.
*While higher order chromatin compaction generally confers greater protection from DSBs, high LET radiation creates non-random damage within these structures, potentially creating more deleterious damage than their decondensed counterparts.

Nucleosomes suppress formation of DSBs during attempted BER of clustered lesions

The close association between DNA and the histone octamer inhibits access by the large radiomimetic compounds described earlier (e.g. (Wu et al., 1999)). As a result, damage from such compounds is confined largely to linker DNA and regions such as promoters, where nucleosome occupancy is lower than in transcriptionally quiescent regions. ROS, on the other hand, are much smaller, and can react with most of the DNA in nucleosomes. Thus, the overall extent to which nucleosomes mask ROS-reactive moieties in DNA is modest, although one commonly sees a ~10 bp periodicity in ROS reactivity, which matches the helical repeat of DNA in nucleosomes (Franchet-Beuzit et al., 1993; Hayes et al., 1990). This periodicity, which is also evident when nucleosomes are probed with DNA endonucleases whose activities are similarly sensitive to groove width (e.g. (Prunell, 1983)), can be attributed to three structural features, acting in concert. First, most DNA segments adopt a discrete helical orientation when bent, either free in space or when wrapped about a histone octamer (Drew and Travers, 1985; Satchwell et al., 1986). Second, moieties on the outside of a bend are typically more reactive than those on the inside because of differences in the width of the minor groove. Finally, reactions between ROS and nucleosomal DNA are impeded by the discrete interactions between histones and the DNA backbone at every helical turn (Luger et al., 1997).

In vitro studies indicate that BER enzymes can repair oxidative lesions at many sites in nucleosomes, without irreversibly disrupting the host nucleosome (reviewed in (Odell et al., 2013)). Thus, BER may differ from other DNA repair pathways, where nucleosome disruption is thought to precede or accompany repair. The efficiency with which single oxidative lesions are repaired varies both with the distance between the lesion and the edge of the nucleosome, and the lesion’s helical orientation relative to the underlying histone octamer ((Odell et al., 2013); also, (Maher et al., 2013; Menoni et al., 2012; Rodriguez and Smerdon, 2013; Ye et al., 2012)). Given that nucleosomes are semi-permissive for BER of single oxidative lesions, one might expect that nucleosomes would offer only limited protection from the BER-dependent conversion of clustered lesions to DSBs. When tested however, nucleosomes suppressed DSB formation at clustered lesions far more than predicted (Cannan et al., 2014). The underlying reasons for this are, first, that the processing of one oxidative lesion will interfere with processing of an opposing strand lesion unless the two lesions are offset from one another by at least three bp (Figure 4). Second, BER of clustered, opposing strand lesions will generate DSBs only if the repair at each lesion is initiated more or less simultaneously and proceeds with similar kinetics. This can only occur in nucleosomes if both opposing strand lesions are sterically accessible. BER factors are able to bind and process sterically occluded lesions in nucleosomes, but only during the relatively infrequent episodes of spontaneous, transient partial unwrapping of DNA from the histone octamer (Maher et al., 2013; Prasad et al., 2007). The upshot is that only a small fraction of the clustered lesions that form in nucleosomes exposed to IR will exhibit a configuration that satisfies both of these geometric constraints.

An external file that holds a picture, illustration, etc.
Object name is nihms810463f4.jpg
Nucleosomes suppress BER-mediated double strand break formation

(A) depicts the excision of an oxidized base (red hexagon) from a nucleosome by a DNA glycosylase (“Gly”). This excision reaction is relatively high efficient when the lesion is oriented so that it can flip through the major groove (red arrow) without steric hindrance from the histone octamer or nearby DNA, into the active site of the glycosylase, which must be able to bind via the minor groove (white arrow). Provided both these constraints are satisfied, base excision repair can proceed to completion, as depicted in Figure 1 (left). If DNA glycosylases initiate repair of two, closely-spaced lesions on opposing strands at about the same time, subsequent steps in BER will generate single strand break or gapped repair intermediates. If these intermediate are present at the same time they may spontaneously convert into a DSB, as described in Figure 1 (right). However, if the opposing strand lesions are separated by fewer than 3 bps, as depicted in (B), near-simultaneous repair cannot occur, probably because processing of one lesion degrades the binding site needed to initiate repair of the second lesion. This restriction is evident in repair reactions with both DNA and nucleosomal substrates. If the opposing strand lesions are more optimally spaced with respect to one another (e.g. 3 or 7 bp), access to one or both lesions may be hindered by the histone octamer, as depicted in (C). In this case, the more accessible lesion will likely be repaired more rapidly than the opposing strand lesion, where repair can begin only when the lesion is exposed by spontaneous, transient partial unwrapping of DNA from the histone octamer (Maher et al., 2013; Prasad et al., 2007). If the opposing strand lesions are optimally spaced (~4–6 bp), and optimally oriented with respect to the underlying histone octamer, as depicted in (D), near-simultaneous BER may ensue, resulting in a DSB (for additional details, see (Cannan et al., 2014)).

The extent to which nucleosomes suppress DSB formation in vitro is substantial, but it is difficult to estimate the magnitude of protection that nucleosomes offer in cells. Dynamic behaviors intrinsic to nucleosomes, such as transient, partial unwrapping of DNA from the histone octamer, will likely reduce their protective effect, as may extrinsic factors such as histone chaperones and chromatin remodeling agents. (The possibility that chromatin-remodeling agents facilitate BER in cells has yet to be convincingly demonstrated.) On the other hand, linker associated factors such as histones H1 and H5, which limit nucleosome mobility (Pennings et al., 1994) and stabilize higher order coiling of the chromatin fiber, probably enhance the nucleosome-mediated protection from DSB formation. This prediction is consistent with studies summarized below, which indicate, collectively, that transcriptionally competent or active euchromatin is more vulnerable to IR than is heterochromatin. It is likely that the elevated vulnerability of euchromatin to DSBs is due in part to more efficient attempted BER of clustered lesions.

The impact of higher order chromatin structure on rates of DSB formation

The distinction between hetero- and euchromatic regions in chromosomes dates to the early 20th century (Heitz, 1928; Heitz, 1929). “Regions of increased gene expression,” or RIDGEs (Caron et al., 2001; Versteeg et al., 2003), reside within euchromatic regions. The chromatin within these regions is in a relatively open configuration (Goetze et al., 2007). Controversy over whether euchromatin is more susceptible to IR-induced DSB formation than heterochromatin persisted for many years, due mainly to differing experimental conditions and methods. Early studies that relied on counting chromosome aberrations in mitotic spreads suggested that heterochromatin is more vulnerable to chromosome breaks than euchromatin (e.g. (Natarajan and Ahnstrom, 1969). However, later studies indicated that translocations occur more frequently in euchromatin (reviewed in (Folle et al., 1998)). By combining pulsed field gel electrophoresis with Southern blotting, (Chiu et al., 1982) were able to show that housekeeping genes are more vulnerable to IR-induced DSBs than α-satellite or bulk DNA. As well, active genes were often found to occur in clusters, surrounded by more radiosensitive chromatin (Bunch et al., 1995). Later studies that used immunostaining of γ-H2AX to monitor DSB formation generally confirmed results from the pulsed field gel electrophoresis studies. Despite the near universal adoption of the immunostaining assays, it is still difficult at times to compare results from different labs. A particularly critical variable in accurately assessing the relative abundance of DSBs in hetero- and euchromatin is the time interval between irradiation and visualization of γ-H2AX foci. If the interval is too brief, foci formation will not be maximal; if too long, a significant portion of the damage may have already been repaired. A related challenge is the apparent suppression of γ-H2AX foci formation in heterochromatin (Kim et al., 2007) and the correspondingly slow repair of DSBs in heterochromatin ((Slijepcevic and Natarajan, 1994); reviewed in (Falk et al., 2010)). Indeed, DSB repair may not occur at all until heterochromatin decondenses (Kruhlak et al., 2006; Lorat et al., 2012). Thus, slow or delayed repair could explain reports of elevated radiation sensitivity in condensed chromatin (Chapman et al., 2001; Stobbe et al., 2002). One of the most compelling studies on the vulnerability of transcriptionally active chromatin to DSB formation combined use of γ-H2AX immunofluorescence with Immuno-FISH, to quantify IR-induced DSBs in individual chromosomes (Falk et al., 2008). The authors compared human chromosome 18, which is largely transcriptionally silent, to chromosome 19, which is similar in size but highly transcribed. Chromosome 19 proved to be ~four-fold more sensitive to IR-induced DSBs than chromosome 18. The authors also compared two 11 Mbp regions (one RIDGE and one ‘anti-RIDGE’ region), both on Chromosome 11. Here too, the RIDGE region was ~four-fold more vulnerable to DSBs than was the anti-RIDGE region. This study strongly suggests that gene density alone can account for a higher than normal sensitivity to IR.

X-ray scattering studies of isolated nuclei have revealed diffraction peaks at 30–40 nm (reflecting side-by-side packing of ~30 nm chromatin filaments), 11 and 6 nm (reflecting the packing constraints of individual nucleosomes), and 2.1 and 2.7 nm (reflecting the diameter of DNA and its pitch within nucleosomes, respectively) ((Langmore and Paulson, 1983), and references therein). This is evidence of relatively uniform DNA packaging at multiple levels. Given this, one might expect chromatin to influence DSB patterns on a scale ranging from ~150 bp (the amount of DNA in one nucleosome) to ~100,000 bp (corresponding to the distance between adjacent chromatin loops in mitotic chromosomes). In asking if IR-induced DSB patterns reflect the packaging of DNA in chromatin, or if higher order chromatin coiling protects DNA from DSB formation, one should also consider whether the damaging agent is low or high LET IR. As noted earlier, low LET IR consists largely of electrons and photons that liberate secondary electrons (β particles) and produce ROS. Although low LET IR can create closely-spaced, clustered lesions (Chatterjee and Holley, 1991), collisions between low LET particles and atoms in tissues are infrequent, and thus low to moderate levels of low-LET may produce DSBs that are more or less randomly distributed. High LET α particles and heavy ions travel along straight tracks and exhibit far higher collision rates, producing DSB patterns that appear to be non-random (Lobrich et al., 1996; Prise et al., 2001; Rydberg, 2001). The distribution of high LET IR-induced damage became more random when chromatin proteins were removed prior to IR (Radulescu et al., 2004). This is direct evidence that non-random DSB patterns from high LET IR reflect the packaging of DNA in chromatin. It also supports the use of IR-induced DSB patterns to determine nucleosome positions and investigate the structure of the 30-nm chromatin fiber (Bernhardt et al., 2003; Rydberg, 2001; Rydberg et al., 1998). A discussion of how these DSB patterns relate to the competing zig-zag and “solenoid” models of higher order chromatin would take us beyond the scope of this review, but interested readers may wish to consult one or more of the recent reviews of the subject (Grigoryev and Woodcock, 2012; Joti et al., 2012; Mozziconacci and Lavelle, 2009; Tremethick, 2007; van Holde and Zlatanova, 2007).

Chromothripsis

In cancers of all kinds, 2–3% of cells contain chromosomes that show signs of having suffered numerous DSBs, inversions, and deletions. When such scars are evident in multiple chromosomes, this phenomenon is referred to as chromoplexy. In chromothripsis (“chromosome shattering”), the chromosome rearrangement and deletion events can number in the hundreds, but are usually confined to a single chromosome (reviewed in (Forment et al., 2012)). Other distinguishing features of chromothripsis include a low gene copy number within shattered regions, and alternations between retained heterozygosity and loss of heterozygosity (Maher and Wilson, 2012)). The chromosome rearrangement and deletion events point to the involvement of NHEJ during the repair of shattered chromosomes, while DNA regions that exhibit loss of heterozygosity may result from break induced DNA replication.

But what generates so many DSBs in the first place? There are several hypotheses, none mutually exclusive. One hypothesis is that chromothripsis reflects an aborted apoptosis (Tubio and Estivill, 2011), where DNA was partially fragmented by endonucleases that are activated during apoptosis. Although reversal of apoptosis has been observed (Tang et al., 2012), this model does not explain why the endonucleolytic activity would have been confined to a single chromosome. A second model posits that NHEJ of eroded telomeres generates sister chromatid fusions. This produces a dicentric chromosome that may be pulled to opposite poles during mitosis. Such events often end when the chromosome breaks, producing new DSBs. This sets the stage for another round of NHEJ that restores the dicentric chromosome, allowing the breakage-fusion cycle to repeat (Colnaghi et al., 2011; Meyerson and Pellman, 2011). In this model, the chromothriptic phenotype develops over the course of many cell cycles. One problem with this theory is that chromosomal termini are not usually lost during chromothripsis. Thus, the postulated initial sister chromosome fusion event would more likely be triggered by a localized set of DSBs than by telomere attrition. A third, and perhaps the most widely accepted, model begins with the mis-segregation of a chromosome during mitosis, which would occur if sister chromatids were slow to decatenate following DNA replication or sister chromatid recombination. As a result, the lagging chromosome may be excluded from the daughter nucleus that reassembles after telophase. Instead, the lagging chromosome is packaged into a separate, small micronucleus that may be starved for DNA replication and repair factors, owing to an insufficient number of nuclear pores. The resulting lag in DNA replication of the chromosome within the micronucleus may not become apparent until nuclei break down in advance of mitosis. At that point, the incompletely replicated chromosome will trigger a DNA damage response, but one that may decay before replication is complete (Giunta et al., 2010; van Vugt et al., 2010). As cells enter metaphase, the incompletely replicated chromosome is “pulverized,” perhaps as a result of premature chromosome condensation. Micronuclei may repeatedly reassemble around chromosomes or chromosome fragments that are no longer capable of normal disjunction. Thus pulverization and rearrangement may occur over multiple cell divisions, until the damaged, rearranged chromosome, or fragments thereof, are taken up by the nucleus (Crasta et al., 2012).

The findings and concepts outlined in this review suggest one additional model that might also contribute to chromothripsis, namely that the numerous, localized DSBs could be generated by collisions between a few high LET particles and a single mitotic chromosome. As noted earlier, a single high LET particle can generate multiple DSBs in relatively small region of the nucleus. Chromatin packaging can bring widely spaced DNA segments into close proximity with one another, increasing the likelihood that both might be damaged by the same high LET particle. Collisions between a high LET particle and a mitotic chromosome may thus produce DSBs separated by hundreds of kilobases. NHEJ of these multiple DSBs during the subsequent G1 phase might then produce the multiple deletions and rearrangements characteristic of chromothripsis. We are attracted to this hypothesis because exposure to a single dose of high energy IR readily fits the descriptions of both non-random damage and a “single catastrophic cellular event.”

Acknowledgments

We thank Drs. Joyce Heckman, Robyn Maher, and Susan Wallace for critical comments on the manuscript.

Work in the Pederson lab is supported in part by NCI grant P01-CA098993.

The abbreviations used are

γ-H2AXphosphorylated histone H2AX
ATMataxia telangiectasia mutated
ATRataxia telangiectasia and Rad3 related
BERbase excision repair
bpbase pair
Cer-XLFCernunnos-X4-like factor
DSBdouble-strand break
HRRhomology-directed, recombination-mediated repair
HZEhigh-Z, high energy
IRionizing radiation
LETlinear energy transfer
MMEJmicrohomology-mediated end-joining
NHEJnon-homologous end joining
PARP1poly-ADP ribose polymerase 1
RIDGEregions of highly expressed genes
ROSreactive oxygen species
SDSAsynthesis-dependent strand annealing
SSBsingle-strand break
XRCC1X-ray repair cross-complementing protein 1
XRCC4X-ray repair cross-complementing protein 4

Footnotes

The authors of this study have no financial conflicts of interest.

References

  • Asaithamby A, Chen DJ. Mechanism of cluster DNA damage repair in response to high-atomic number and energy particles radiation. Mutation research. 2011;711(1–2):87–99. [PMC free article] [PubMed] [Google Scholar]
  • Averbeck NB, Ringel O, Herrlitz M, Jakob B, Durante M, Taucher-Scholz G. DNA end resection is needed for the repair of complex lesions in G1-phase human cells. Cell Cycle. 2014;13(16):2509–2516. [PMC free article] [PubMed] [Google Scholar]
  • Baker DN, Jaynes AN, Hoxie VC, Thorne RM, Foster JC, Li X, Fennell JF, Wygant JR, Kanekal SG, Erickson PJ, Kurth W, Li W, Ma Q, Schiller Q, Blum L, Malaspina DM, Gerrard A, Lanzerotti LJ. An impenetrable barrier to ultrarelativistic electrons in the Van Allen radiation belts. Nature. 2014;515(7528):531–534. [PubMed] [Google Scholar]
  • Balasubramanian B, Pogozelski WK, Tullius TD. DNA strand breaking by the hydroxyl radical is governed by the accessible surface areas of the hydrogen atoms of the DNA backbone. Proceedings of the National Academy of Sciences of the United States of America. 1998;95(17):9738–9743. [PMC free article] [PubMed] [Google Scholar]
  • Barnard S, Bouffler S, Rothkamm K. The shape of the radiation dose response for DNA double-strand break induction and repair. Genome integrity. 2013;4(1):1. [PMC free article] [PubMed] [Google Scholar]
  • Bernhardt P, Friedland W, Jacob P, Paretzke H. Modeling of ultrasoft X-ray induced DNA damage using structured higher order DNA targets. International Journal of Mass Spectrometry. 2003;223:579–597. [Google Scholar]
  • Blaisdell JO, Harrison L, Wallace SS. Base excision repair processing of radiation-induced clustered DNA lesions. Radiat Prot Dosimetry. 2001;97(1):25–31. [PubMed] [Google Scholar]
  • Blaisdell JO, Wallace SS. Abortive base-excision repair of radiation-induced clustered DNA lesions in Escherichia coli. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(13):7426–7430. [PMC free article] [PubMed] [Google Scholar]
  • Bunch RT, Gewirtz DA, Povirk LF. Ionizing radiation-induced DNA strand breakage and rejoining in specific genomic regions as determined by an alkaline unwinding/Southern blotting method. Int J Radiat Biol. 1995;68(5):553–562. [PubMed] [Google Scholar]
  • Cannan WJ, Tsang BP, Wallace SS, Pederson DS. Nucleosomes suppress the formation of double-strand DNA breaks during attempted base excision repair of clustered oxidative damages. The Journal of biological chemistry. 2014;289(29):19881–19893. [PMC free article] [PubMed] [Google Scholar]
  • Caron H, van Schaik B, van der Mee M, Baas F, Riggins G, van Sluis P, Hermus MC, van Asperen R, Boon K, Voute PA, Heisterkamp S, van Kampen A, Versteeg R. Science. 5507. Vol. 291. New York, NY: 2001. The human transcriptome map: clustering of highly expressed genes in chromosomal domains; pp. 1289–1292. [PubMed] [Google Scholar]
  • Chapman JD, Stobbe CC, Matsumoto Y. Chromatin compaction and tumor cell radiosensitivity at 2 gray. American journal of clinical oncology. 2001;24(5):509–515. [PubMed] [Google Scholar]
  • Chatterjee A, Holley WR. Energy deposition mechanisms and biochemical aspects of DNA strand breaks by ionizing radiation. International journal of quantum chemistry. 1991;39(5):709–727. [Google Scholar]
  • Chen J, Stubbe J. Bleomycins: towards better therapeutics. Nat Rev Cancer. 2005;5(2):102–112. [PubMed] [Google Scholar]
  • Chiu S, Oleinick NL. Radioprotection of cellular chromatin by the polyamines spermine and putrescine: preferential action against formation of DNA-protein crosslinks. Radiat Res. 1998;149(6):543–549. [PubMed] [Google Scholar]
  • Chiu SM, Oleinick NL, Friedman LR, Stambrook PJ. Hypersensitivity of DNA in transcriptionally active chromatin to ionizing radiation. Biochimica et biophysica acta. 1982;699(1):15–21. [PubMed] [Google Scholar]
  • Colnaghi R, Carpenter G, Volker M, O'Driscoll M. The consequences of structural genomic alterations in humans: genomic disorders, genomic instability and cancer. Semin Cell Dev Biol. 2011;22(8):875–885. [PubMed] [Google Scholar]
  • Crasta K, Ganem NJ, Dagher R, Lantermann AB, Ivanova EV, Pan Y, Nezi L, Protopopov A, Chowdhury D, Pellman D. DNA breaks and chromosome pulverization from errors in mitosis. Nature. 2012;482(7383):53–58. [PMC free article] [PubMed] [Google Scholar]
  • Dahm-Daphi J, Sass C, Alberti W. Comparison of biological effects of DNA damage induced by ionizing radiation and hydrogen peroxide in CHO cells. Int J Radiat Biol. 2000;76(1):67–75. [PubMed] [Google Scholar]
  • David SS, O'Shea VL, Kundu S. Base-excision repair of oxidative DNA damage. Nature. 2007;447(7147):941–950. [PMC free article] [PubMed] [Google Scholar]
  • De Bont R, van Larebeke N. Endogenous DNA damage in humans: a review of quantitative data. Mutagenesis. 2004;19(3):169–185. [PubMed] [Google Scholar]
  • de Massy B. Initiation of meiotic recombination: how and where? Conservation and specificities among eukaryotes. Annual review of genetics. 2013;47:563–599. [PubMed] [Google Scholar]
  • Dextraze ME, Gantchev T, Girouard S, Hunting D. DNA interstrand cross-links induced by ionizing radiation: an unsung lesion. Mutation research. 2010;704(1–3):101–107. [PubMed] [Google Scholar]
  • Doss M, Little MP, Orton CG. Point/Counterpoint: low-dose radiation is beneficial, not harmful. Medical physics. 2014;41(7):070601. [PMC free article] [PubMed] [Google Scholar]
  • Drew HR, Travers AA. DNA bending and its relation to nucleosome positioning. Journal of molecular biology. 1985;186(4):773–790. [PubMed] [Google Scholar]
  • Duclos S, Doublié S, Wallace SS. The cellular response to the genotoxic insult: The question of threshold for genotoxic carcinogens. Cambridge, UK: The Royal Society of Chemistry; 2012. Consequences and repair of oxidative DNA damage; pp. 115–159. [Google Scholar]
  • Duharcourt S, Lepere G, Meyer E. Developmental genome rearrangements in ciliates: a natural genomic subtraction mediated by non-coding transcripts. Trends Genet. 2009;25(8):344–350. [PubMed] [Google Scholar]
  • Eccles LJ, O'Neill P, Lomax ME. Delayed repair of radiation induced clustered DNA damage: friend or foe? Mutation research. 2011;711(1–2):134–141. [PMC free article] [PubMed] [Google Scholar]
  • Einstein A. Über einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt. Annalen der Physik. 1905;322(6):132–148. [Google Scholar]
  • Falk M, Lukasova E, Kozubek S. Chromatin structure influences the sensitivity of DNA to gamma-radiation. Biochimica et biophysica acta. 2008;1783(12):2398–2414. [PubMed] [Google Scholar]
  • Falk M, Lukasova E, Kozubek S. Higher-order chromatin structure in DSB induction, repair and misrepair. Mutation research. 2010;704(1–3):88–100. [PubMed] [Google Scholar]
  • Finch JT, Klug A. Solenoidal model for superstructure in chromatin. Proceedings of the National Academy of Sciences of the United States of America. 1976;73(6):1897–1901. [PMC free article] [PubMed] [Google Scholar]
  • Folle GA, Martinez-Lopez W, Boccardo E, Obe G. Localization of chromosome breakpoints: implication of the chromatin structure and nuclear architecture. Mutation research. 1998;404(1–2):17–26. [PubMed] [Google Scholar]
  • Forment JV, Kaidi A, Jackson SP. Chromothripsis and cancer: causes and consequences of chromosome shattering. Nat Rev Cancer. 2012;12(10):663–670. [PubMed] [Google Scholar]
  • Fraga CG, Shigenaga MK, Park JW, Degan P, Ames BN. Oxidative damage to DNA during aging: 8-hydroxy-2'-deoxyguanosine in rat organ DNA and urine. Proceedings of the National Academy of Sciences of the United States of America. 1990;87(12):4533–4537. [PMC free article] [PubMed] [Google Scholar]
  • Franchet-Beuzit J, Spotheim-Maurizot M, Sabattier R, Blazy-Baudras B, Charlier M. Radiolytic footprinting. Beta rays, gamma photons, and fast neutrons probe DNA-protein interactions. Biochemistry. 1993;32(8):2104–2110. [PubMed] [Google Scholar]
  • Frankenberg-Schwager M, Gebauer A, Koppe C, Wolf H, Pralle E, Frankenberg D. Single-strand annealing, conservative homologous recombination, nonhomologous DNA end joining, and the cell cycle-dependent repair of DNA double-strand breaks induced by sparsely or densely ionizing radiation. Radiat Res. 2009;171(3):265–273. [PubMed] [Google Scholar]
  • Friedberg EC. DNA repair and mutagenesis. Washington D.C.: ASM Press; 2006. [Google Scholar]
  • Giunta S, Belotserkovskaya R, Jackson SP. DNA damage signaling in response to double-strand breaks during mitosis. The Journal of cell biology. 2010;190(2):197–207. [PMC free article] [PubMed] [Google Scholar]
  • Goetze S, Mateos-Langerak J, Gierman HJ, de Leeuw W, Giromus O, Indemans MH, Koster J, Ondrej V, Versteeg R, van Driel R. The three-dimensional structure of human interphase chromosomes is related to the transcriptome map. Molecular and cellular biology. 2007;27(12):4475–4487. [PMC free article] [PubMed] [Google Scholar]
  • Goodhead DT, Thacker J, Cox R. Weiss Lecture. Effects of radiations of different qualities on cells: molecular mechanisms of damage and repair. Int J Radiat Biol. 1993;63(5):543–556. [PubMed] [Google Scholar]
  • Grigoryev SA, Woodcock CL. Chromatin organization - the 30 nm fiber. Exp Cell Res. 2012;318(12):1448–1455. [PubMed] [Google Scholar]
  • Haber JE. Mating-type genes and MAT switching in Saccharomyces cerevisiae. Genetics. 2012;191(1):33–64. [PMC free article] [PubMed] [Google Scholar]
  • Hamperl S, Cimprich KA. The contribution of co-transcriptional RNA:DNA hybrid structures to DNA damage and genome instability. DNA repair. 2014;19:84–94. [PMC free article] [PubMed] [Google Scholar]
  • Hande KR. Etoposide: four decades of development of a topoisomerase II inhibitor. European Journal of Cancer. 1998;34(10):1514–1521. [PubMed] [Google Scholar]
  • Hayes JJ, Tullius TD, Wolffe AP. The structure of DNA in a nucleosome. Proceedings of the National Academy of Sciences of the United States of America. 1990;87(19):7405–7409. [PMC free article] [PubMed] [Google Scholar]
  • Hegde ML, Hazra TK, Mitra S. Early steps in the DNA base excision/single-strand interruption repair pathway in mammalian cells. Cell research. 2008;18(1):27–47. [PMC free article] [PubMed] [Google Scholar]
  • Heitz E. Das Heterochromatin. der Moose: Bornträger; 1928. [Google Scholar]
  • Heitz E. Heterochromatin, chromocentren, chromomenen. Ber Deutsch Bot Ges. 1929;47:274–284. [Google Scholar]
  • Hinton JA, Hofmann W. Teraelectronvolt Astronomy. Annual Review of Astronomy and Astrophysics. 2009;47(1):523–565. [Google Scholar]
  • Huertas P, Cortes-Ledesma F, Sartori AA, Aguilera A, Jackson SP. CDK targets Sae2 to control DNA-end resection and homologous recombination. Nature. 2008;455(7213):689–692. [PMC free article] [PubMed] [Google Scholar]
  • Isabelle V, Prevost C, Spotheim-Maurizot M, Sabattier R, Charlier M. Radiation-induced damages in single- and double-stranded DNA. Int J Radiat Biol. 1995;67(2):169–176. [PubMed] [Google Scholar]
  • Joti Y, Hikima T, Nishino Y, Kamada F, Hihara S, Takata H, Ishikawa T, Maeshima K. Chromosomes without a 30-nm chromatin fiber. Nucleus. 2012;3(5):404–410. [PMC free article] [PubMed] [Google Scholar]
  • Kim JA, Kruhlak M, Dotiwala F, Nussenzweig A, Haber JE. Heterochromatin is refractory to gamma-H2AX modification in yeast and mammals. The Journal of cell biology. 2007;178(2):209–218. [PMC free article] [PubMed] [Google Scholar]
  • Kim M-HY, Wilson JW, Cucinotta FA, Simonsen LC, Atwell W, Badavi FF, Miller J. Contribution of high charge and energy (HZE) ions during solar-particle event of September 29, 1989. 1999 [Google Scholar]
  • Koster DA, Palle K, Bot ES, Bjornsti MA, Dekker NH. Antitumour drugs impede DNA uncoiling by topoisomerase I. Nature. 2007;448(7150):213–217. [PubMed] [Google Scholar]
  • Krokan HE, Bjoras M. Base excision repair. Cold Spring Harb Perspect Biol. 2013;5(4):a012583. [PMC free article] [PubMed] [Google Scholar]
  • Kruhlak MJ, Celeste A, Dellaire G, Fernandez-Capetillo O, Muller WG, McNally JG, Bazett-Jones DP, Nussenzweig A. Changes in chromatin structure and mobility in living cells at sites of DNA double-strand breaks. The Journal of cell biology. 2006;172(6):823–834. [PMC free article] [PubMed] [Google Scholar]
  • Langmore JP, Paulson JR. Low angle x-ray diffraction studies of chromatin structure in vivo and in isolated nuclei and metaphase chromosomes. The Journal of cell biology. 1983;96(4):1120–1131. [PMC free article] [PubMed] [Google Scholar]
  • Lavelle C, Foray N. Chromatin structure and radiation-induced DNA damage: from structural biology to radiobiology. The international journal of biochemistry & cell biology. 2014;49:84–97. [PubMed] [Google Scholar]
  • Lieber MR. The mechanism of double-strand DNA break repair by the nonhomologous DNA end-joining pathway. Annual review of biochemistry. 2010;79:181–211. [PMC free article] [PubMed] [Google Scholar]
  • Lindahl T, Nyberg B. Rate of depurination of native deoxyribonucleic acid. Biochemistry. 1972;11(19):3610–3618. [PubMed] [Google Scholar]
  • Little MP, Azizova TV, Bazyka D, Bouffler SD, Cardis E, Chekin S, Chumak VV, Cucinotta FA, de Vathaire F, Hall P, Harrison JD, Hildebrandt G, Ivanov V, Kashcheev VV, Klymenko SV, Kreuzer M, Laurent O, Ozasa K, Schneider T, Tapio S, Taylor AM, Tzoulaki I, Vandoolaeghe WL, Wakeford R, Zablotska LB, Zhang W, Lipshultz SE. Systematic review and meta-analysis of circulatory disease from exposure to low-level ionizing radiation and estimates of potential population mortality risks. Environ Health Perspect. 2012;120(11):1503–1511. [PMC free article] [PubMed] [Google Scholar]
  • Liu LF, Wang JC. Supercoiling of the DNA template during transcription. Proceedings of the National Academy of Sciences of the United States of America. 1987;84(20):7024–7027. [PMC free article] [PubMed] [Google Scholar]
  • Lobrich M, Cooper PK, Rydberg B. Non-random distribution of DNA double-strand breaks induced by particle irradiation. Int J Radiat Biol. 1996;70(5):493–503. [PubMed] [Google Scholar]
  • Lorat Y, Schanz S, Schuler N, Wennemuth G, Rube C, Rube CE. Beyond repair foci: DNA double-strand break repair in euchromatic and heterochromatic compartments analyzed by transmission electron microscopy. PloS one. 2012;7(5):e38165. [PMC free article] [PubMed] [Google Scholar]
  • Luger K, Mader AW, Richmond RK, Sargent DF, Richmond TJ. Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature. 1997;389(6648):251–260. [PubMed] [Google Scholar]
  • Maher CA, Wilson RK. Chromothripsis and human disease: piecing together the shattering process. Cell. 2012;148(1–2):29–32. [PMC free article] [PubMed] [Google Scholar]
  • Maher RL, Prasad A, Rizvanova O, Wallace SS, Pederson DS. Contribution of DNA unwrapping from histone octamers to the repair of oxidatively damaged DNA in nucleosomes. DNA repair. 2013;12(11):964–971. [PMC free article] [PubMed] [Google Scholar]
  • Malkova A, Haber JE. Mutations arising during repair of chromosome breaks. Annual review of genetics. 2012;46:455–473. [PubMed] [Google Scholar]
  • Malkova A, Ira G. Break-induced replication: functions and molecular mechanism. Current opinion in genetics & development. 2013;23(3):271–279. [PMC free article] [PubMed] [Google Scholar]
  • Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER, 3rd, Hurov KE, Luo J, Bakalarski CE, Zhao Z, Solimini N, Lerenthal Y, Shiloh Y, Gygi SP, Elledge SJ. Science. 5828. Vol. 316. New York, NY: 2007. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage; pp. 1160–1166. [PubMed] [Google Scholar]
  • Mehta A, Haber JE. Sources of DNA double-strand breaks and models of recombinational DNA repair. Cold Spring Harb Perspect Biol. 2014;6(9):a016428. [PMC free article] [PubMed] [Google Scholar]
  • Menoni H, Shukla MS, Gerson V, Dimitrov S, Angelov D. Base excision repair of 8-oxoG in dinucleosomes. Nucleic acids research. 2012;40(2):692–700. [PMC free article] [PubMed] [Google Scholar]
  • Mewaldt RA. Galactic cosmic ray composition and energy spectra. Advances in space research : the official journal of the Committee on Space Research. 1994;14(10):737–747. [PubMed] [Google Scholar]
  • Meyerson M, Pellman D. Cancer genomes evolve by pulverizing single chromosomes. Cell. 2011;144(1):9–10. [PubMed] [Google Scholar]
  • Mimitou EP, Symington LS. Nucleases and helicases take center stage in homologous recombination. Trends in biochemical sciences. 2009;34(5):264–272. [PubMed] [Google Scholar]
  • Mirkin EV, Mirkin SM. Replication fork stalling at natural impediments. Microbiol Mol Biol Rev. 2007;71(1):13–35. [PMC free article] [PubMed] [Google Scholar]
  • Mozziconacci J, Lavelle C. Chromatin Fiber: 30 years of models. Computational Biology: New Research. 2009:147–163. [Google Scholar]
  • NASA. Strategic program plan for space radiation health research. Washington, DC: NASA; 1998. [Google Scholar]
  • NASA. APOD: 2006 August 14 - Cosmic Rays. 2006 [Google Scholar]
  • NASA. Phillips T. Who's Afraid of a Solar Flare? 2005 [Google Scholar]
  • Natarajan AT, Ahnstrom G. Heterochromatin and chromosome aberrations. Chromosoma. 1969;28(1):48–61. [PubMed] [Google Scholar]
  • Odell ID, Wallace SS, Pederson DS. Rules of engagement for base excision repair in chromatin. J Cell Physiol. 2013;228(2):258–266. [PMC free article] [PubMed] [Google Scholar]
  • Olins AL, Olins DE. Science. 4122. Vol. 183. New York, NY: 1974. Spheroid chromatin units (v bodies) pp. 330–332. [PubMed] [Google Scholar]
  • Olins AL, Olins DE. Stereo electron microscopy of the 25-nm chromatin fibers in isolated nuclei. The Journal of cell biology. 1979;81(1):260–265. [PMC free article] [PubMed] [Google Scholar]
  • Oudet P, Gross-Bellard M, Chambon P. Electron microscopic and biochemical evidence that chromatin structure is a repeating unit. Cell. 1975;4(4):281–300. [PubMed] [Google Scholar]
  • Pennings S, Meersseman G, Bradbury EM. Linker histones H1 and H5 prevent the mobility of positioned nucleosomes. Proceedings of the National Academy of Sciences of the United States of America. 1994;91(22):10275–10279. [PMC free article] [PubMed] [Google Scholar]
  • Pernot E, Hall J, Baatout S, Benotmane MA, Blanchardon E, Bouffler S, El Saghire H, Gomolka M, Guertler A, Harms-Ringdahl M, Jeggo P, Kreuzer M, Laurier D, Lindholm C, Mkacher R, Quintens R, Rothkamm K, Sabatier L, Tapio S, de Vathaire F, Cardis E. Ionizing radiation biomarkers for potential use in epidemiological studies. Mutation research. 2012;751(2):258–286. [PubMed] [Google Scholar]
  • Peterson SE, Li Y, Wu-Baer F, Chait BT, Baer R, Yan H, Gottesman ME, Gautier J. Activation of DSB processing requires phosphorylation of CtIP by ATR. Molecular cell. 2013;49(4):657–667. [PMC free article] [PubMed] [Google Scholar]
  • Pfeiffer P, Goedecke W, Obe G. Mechanisms of DNA double-strand break repair and their potential to induce chromosomal aberrations. Mutagenesis. 2000;15(4):289–302. [PubMed] [Google Scholar]
  • Pierre Auger Observatory. The Mystery of High-Energy Cosmic Rays [Google Scholar]
  • Pinto M, Prise KM, Michael BD. Evidence for complexity at the nanometer scale of radiation-induced DNA DSBs as a determinant of rejoining kinetics. Radiat Res. 2005;164(1):73–85. [PubMed] [Google Scholar]
  • Polo SE, Jackson SP. Dynamics of DNA damage response proteins at DNA breaks: a focus on protein modifications. Genes & development. 2011;25(5):409–433. [PMC free article] [PubMed] [Google Scholar]
  • Pooley AS, Pardon JF, Richards BM. The relation between the unit thread of chromosomes and isolated nucleohistone. Journal of molecular biology. 1974;85(4):533–549. [PubMed] [Google Scholar]
  • Povirk LF. DNA damage and mutagenesis by radiomimetic DNA-cleaving agents: bleomycin, neocarzinostatin and other enediynes. Mutation research. 1996;355(1–2):71–89. [PubMed] [Google Scholar]
  • Prado F, Aguilera A. Impairment of replication fork progression mediates RNA polII transcription-associated recombination. The EMBO journal. 2005;24(6):1267–1276. [PMC free article] [PubMed] [Google Scholar]
  • Prasad A, Wallace SS, Pederson DS. Initiation of base excision repair of oxidative lesions in nucleosomes by the human, bifunctional DNA glycosylase NTH1. Molecular and cellular biology. 2007;27(24):8442–8453. [PMC free article] [PubMed] [Google Scholar]
  • Prise KM, Pinto M, Newman HC, Michael BD. A review of studies of ionizing radiation-induced double-strand break clustering. Radiat Res. 2001;156(5 Pt 2):572–576. [PubMed] [Google Scholar]
  • Prunell A. Periodicity of exonuclease III digestion of chromatin and the pitch of deoxyribonucleic acid on the nucleosome. Biochemistry. 1983;22(21):4887–4894. [PubMed] [Google Scholar]
  • Radulescu I, Elmroth K, Stenerlow B. Chromatin organization contributes to non-randomly distributed double-strand breaks after exposure to high-LET radiation. Radiat Res. 2004;161(1):1–8. [PubMed] [Google Scholar]
  • Reynolds P, Anderson JA, Harper JV, Hill MA, Botchway SW, Parker AW, O'Neill P. The dynamics of Ku70/80 and DNA-PKcs at DSBs induced by ionizing radiation is dependent on the complexity of damage. Nucleic acids research. 2012;40(21):10821–10831. [PMC free article] [PubMed] [Google Scholar]
  • Ris H, Kubai DF. Chromosome structure. Annual review of genetics. 1970;4:263–294. [PubMed] [Google Scholar]
  • Robertson AB, Klungland A, Rognes T, Leiros I. DNA repair in mammalian cells: Base excision repair: the long and short of it. Cell Mol Life Sci. 2009;66(6):981–993. [PMC free article] [PubMed] [Google Scholar]
  • Rodriguez Y, Smerdon MJ. The structural location of DNA lesions in nucleosome core particles determines accessibility by base excision repair enzymes. The Journal of biological chemistry. 2013;288(19):13863–13875. [PMC free article] [PubMed] [Google Scholar]
  • Rydberg B. Clusters of DNA damage induced by ionizing radiation: formation of short DNA fragments. II. Experimental detection. Radiat Res. 1996;145(2):200–209. [PubMed] [Google Scholar]
  • Rydberg B. Radiation-induced DNA damage and chromatin structure. Acta oncologica. 2001;40(6):682–685. [PubMed] [Google Scholar]
  • Rydberg B, Holley WR, Mian IS, Chatterjee A. Chromatin conformation in living cells: support for a zig-zag model of the 30 nm chromatin fiber. Journal of molecular biology. 1998;284(1):71–84. [PubMed] [Google Scholar]
  • Sage E, Harrison L. Clustered DNA lesion repair in eukaryotes: relevance to mutagenesis and cell survival. Mutation research. 2011;711(1–2):123–133. [PMC free article] [PubMed] [Google Scholar]
  • San Filippo J, Sung P, Klein H. Mechanism of eukaryotic homologous recombination. Annual review of biochemistry. 2008;77:229–257. [PubMed] [Google Scholar]
  • Satchwell SC, Drew HR, Travers AA. Sequence periodicities in chicken nucleosome core DNA. Journal of molecular biology. 1986;191(4):659–675. [PubMed] [Google Scholar]
  • Sax K. Chromosome Aberrations Induced by X-Rays. Genetics. 1938;23(5):494–516. [PMC free article] [PubMed] [Google Scholar]
  • Semenenko VA, Stewart RD. A fast Monte Carlo algorithm to simulate the spectrum of DNA damages formed by ionizing radiation. Radiat Res. 2004;161(4):451–457. [PubMed] [Google Scholar]
  • Slijepcevic P, Natarajan AT. Distribution of X-ray-induced G2 chromatid damage among Chinese hamster chromosomes: influence of chromatin conformation. Mutation research. 1994;323(3):113–119. [PubMed] [Google Scholar]
  • Smolka MB, Albuquerque CP, Chen SH, Zhou H. Proteome-wide identification of in vivo targets of DNA damage checkpoint kinases. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(25):10364–10369. [PMC free article] [PubMed] [Google Scholar]
  • Sonnenblick A, de Azambuja E, Azim HA, Jr, Piccart M. An update on PARP inhibitors-moving to the adjuvant setting. Nature reviews Clinical oncology. 2015;12(1):27–41. [PubMed] [Google Scholar]
  • Soulas-Sprauel P, Rivera-Munoz P, Malivert L, Le Guyader G, Abramowski V, Revy P, de Villartay JP. V(D)J and immunoglobulin class switch recombinations: a paradigm to study the regulation of DNA end-joining. Oncogene. 2007;26(56):7780–7791. [PubMed] [Google Scholar]
  • Spotheim-Maurizot M, Ruiz S, Sabattier R, Charlier M. Radioprotection of DNA by polyamines. Int J Radiat Biol. 1995;68(5):571–577. [PubMed] [Google Scholar]
  • Staaf E, Brehwens K, Haghdoost S, Czub J, Wojcik A. Gamma-H2AX foci in cells exposed to a mixed beam of X-rays and alpha particles. Genome integrity. 2012;3(1):8. [PMC free article] [PubMed] [Google Scholar]
  • Stavnezer J. Complex regulation and function of activation-induced cytidine deaminase. Trends Immunol. 2011;32(5):194–201. [PMC free article] [PubMed] [Google Scholar]
  • Stobbe CC, Park SJ, Chapman JD. The radiation hypersensitivity of cells at mitosis. Int J Radiat Biol. 2002;78(12):1149–1157. [PubMed] [Google Scholar]
  • Sung W, Park JM, Choi CH, Ha SW, Ye SJ. The effect of photon energy on intensity-modulated radiation therapy (IMRT) plans for prostate cancer. Radiation oncology journal. 2012;30(1):27–35. [PMC free article] [PubMed] [Google Scholar]
  • Sutherland BM, Bennett PV, Sidorkina O, Laval J. Clustered damages and total lesions induced in DNA by ionizing radiation: oxidized bases and strand breaks. Biochemistry. 2000a;39(27):8026–8031. [PubMed] [Google Scholar]
  • Sutherland BM, Bennett PV, Sidorkina O, Laval J. Clustered DNA damages induced in isolated DNA and in human cells by low doses of ionizing radiation. Proceedings of the National Academy of Sciences of the United States of America. 2000b;97(1):103–108. [PMC free article] [PubMed] [Google Scholar]
  • Sutherland BM, Georgakilas AG, Bennett PV, Laval J, Sutherland JC. Quantifying clustered DNA damage induction and repair by gel electrophoresis, electronic imaging and number average length analysis. Mutation research. 2003;531(1–2):93–107. [PubMed] [Google Scholar]
  • Swenberg CE, Speicher JM. Neutron and gamma-radiation sensitivity of plasmid DNA of varying superhelical density. Radiat Res. 1995;144(3):301–309. [PubMed] [Google Scholar]
  • Sy D, Savoye C, Begusova M, Michalik V, Charlier M, Spotheim-Maurizot M. Sequence-dependent variations of DNA structure modulate radiation-induced strand breakage. Int J Radiat Biol. 1997;72(2):147–155. [PubMed] [Google Scholar]
  • Syeda AH, Hawkins M, McGlynn P. Recombination and replication. Cold Spring Harb Perspect Biol. 2014;6(11):a016550. [PMC free article] [PubMed] [Google Scholar]
  • Takahashi T, Burguiere-Slezak G, Van der Kemp PA, Boiteux S. Topoisomerase 1 provokes the formation of short deletions in repeated sequences upon high transcription in Saccharomyces cerevisiae. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(2):692–697. [PMC free article] [PubMed] [Google Scholar]
  • Tang HL, Tang HM, Mak KH, Hu S, Wang SS, Wong KM, Wong CS, Wu HY, Law HT, Liu K, Talbot CC, Jr, Lau WK, Montell DJ, Fung MC. Cell survival, DNA damage, and oncogenic transformation after a transient and reversible apoptotic response. Molecular biology of the cell. 2012;23(12):2240–2252. [PMC free article] [PubMed] [Google Scholar]
  • Terasawa T, Dvorak T, Ip S, Raman G, Lau J, Trikalinos TA. Systematic review: charged-particle radiation therapy for cancer. Ann Intern Med. 2009;151(8):556–565. [PubMed] [Google Scholar]
  • Thompson LH. Recognition, signaling, and repair of DNA double-strand breaks produced by ionizing radiation in mammalian cells: the molecular choreography. Mutation research. 2012;751(2):158–246. [PubMed] [Google Scholar]
  • Tice R, Setlow R. DNA repair and replication in aging organisms and cells. In: Finch CE, Schneider EL, Adelman RC, Martin GM, Masoro EJ, editors. Handbook of the biology of aging. 1985. assistance of associate editors. [Google Scholar]
  • Tichy ED, Pillai R, Deng L, Liang L, Tischfield J, Schwemberger SJ, Babcock GF, Stambrook PJ. Mouse embryonic stem cells, but not somatic cells, predominantly use homologous recombination to repair double-strand DNA breaks. Stem cells and development. 2010;19(11):1699–1711. [PMC free article] [PubMed] [Google Scholar]
  • Tremethick DJ. Higher-order structures of chromatin: the elusive 30 nm fiber. Cell. 2007;128(4):651–654. [PubMed] [Google Scholar]
  • Truong LN, Li Y, Shi LZ, Hwang PY, He J, Wang H, Razavian N, Berns MW, Wu X. Microhomology-mediated End Joining and Homologous Recombination share the initial end resection step to repair DNA double-strand breaks in mammalian cells. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(19):7720–7725. [PMC free article] [PubMed] [Google Scholar]
  • Tubio JM, Estivill X. Cancer: When catastrophe strikes a cell. Nature. 2011;470(7335):476–477. [PubMed] [Google Scholar]
  • Vaidya A, Mao Z, Tian X, Spencer B, Seluanov A, Gorbunova V. Knock-in reporter mice demonstrate that DNA repair by non-homologous end joining declines with age. PLoS Genet. 2014;10(7):e1004511. [PMC free article] [PubMed] [Google Scholar]
  • van Dam L, Korolev N, Nordenskiold L. Polyamine-nucleic acid interactions and the effects on structure in oriented DNA fibers. Nucleic acids research. 2002;30(2):419–428. [PMC free article] [PubMed] [Google Scholar]
  • van Holde K, Zlatanova J. Chromatin fiber structure: Where is the problem now? Semin Cell Dev Biol. 2007;18(5):651–658. [PubMed] [Google Scholar]
  • van Vugt MA, Gardino AK, Linding R, Ostheimer GJ, Reinhardt HC, Ong SE, Tan CS, Miao H, Keezer SM, Li J, Pawson T, Lewis TA, Carr SA, Smerdon SJ, Brummelkamp TR, Yaffe MB. A mitotic phosphorylation feedback network connects Cdk1, Plk1, 53BP1, and Chk2 to inactivate the G(2)/M DNA damage checkpoint. PLoS Biol. 2010;8(1):e1000287. [PMC free article] [PubMed] [Google Scholar]
  • Versteeg R, van Schaik BD, van Batenburg MF, Roos M, Monajemi R, Caron H, Bussemaker HJ, van Kampen AH. The human transcriptome map reveals extremes in gene density, intron length, GC content, and repeat pattern for domains of highly and weakly expressed genes. Genome Res. 2003;13(9):1998–2004. [PMC free article] [PubMed] [Google Scholar]
  • Vilenchik MM, Knudson AG. Endogenous DNA double-strand breaks: production, fidelity of repair, and induction of cancer. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(22):12871–12876. [PMC free article] [PubMed] [Google Scholar]
  • Wahba L, Gore SK, Koshland D. The homologous recombination machinery modulates the formation of RNA-DNA hybrids and associated chromosome instability. eLife. 2013;2:e00505. [PMC free article] [PubMed] [Google Scholar]
  • Walker JR, Corpina RA, Goldberg J. Structure of the Ku heterodimer bound to DNA and its implications for double-strand break repair. Nature. 2001;412(6847):607–614. [PubMed] [Google Scholar]
  • Wallace SS. Base excision repair: a critical player in many games. DNA repair. 2014;19:14–26. [PMC free article] [PubMed] [Google Scholar]
  • Ward J. DNA damage produced by ionizing radiation in mammalian cells: identities, mechanisms of formation, and reparability. Progress in nucleic acid research and molecular biology. 1988;35:95–125. [PubMed] [Google Scholar]
  • Warters RL, Newton GL, Olive PL, Fahey RC. Radioprotection of human cell nuclear DNA by polyamines: radiosensitivity of chromatin is influenced by tightly bound spermine. Radiat Res. 1999;151(3):354–362. [PubMed] [Google Scholar]
  • Weinfeld M, Soderlind KJ. 32P-postlabeling detection of radiation-induced DNA damage: identification and estimation of thymine glycols and phosphoglycolate termini. Biochemistry. 1991;30(4):1091–1097. [PubMed] [Google Scholar]
  • Wellinger RE, Prado F, Aguilera A. Replication fork progression is impaired by transcription in hyperrecombinant yeast cells lacking a functional THO complex. Molecular and cellular biology. 2006;26(8):3327–3334. [PMC free article] [PubMed] [Google Scholar]
  • Woodcock CL, Safer JP, Stanchfield JE. Structural repeating units in chromatin. I. Evidence for their general occurrence. Exp Cell Res. 1976;97:101–110. [PubMed] [Google Scholar]
  • Worcel A, Benyajati C. Higher order coiling of DNA in chromatin. Cell. 1977;12(1):83–100. [PubMed] [Google Scholar]
  • World Health Organization. WHO Handbook on Indoor Radon: A Public Health Perspective. Geneva: 2009. WHO Handbook on Indoor Radon: A Public Health Perspective. [PubMed] [Google Scholar]
  • Wu J, Xu J, Dedon PC. Modulation of enediyne-induced DNA damage by chromatin structures in transcriptionally active genes. Biochemistry. 1999;38(47):15641–15646. [PubMed] [Google Scholar]
  • Yajima H, Fujisawa H, Nakajima NI, Hirakawa H, Jeggo PA, Okayasu R, Fujimori A. The complexity of DNA double strand breaks is a critical factor enhancing end-resection. DNA repair. 2013;12(11):936–946. [PubMed] [Google Scholar]
  • Yamaguchi H, Uchihori Y, Yasuda N, Takada M, Kitamura H. Estimation of yields of OH radicals in water irradiated by ionizing radiation. J Radiat Res. 2005;46(3):333–341. [PubMed] [Google Scholar]
  • Yang N, Chaudhry MA, Wallace SS. Base excision repair by hNTH1 and hOGG1: a two edged sword in the processing of DNA damage in gamma-irradiated human cells. DNA repair. 2006;5(1):43–51. [PubMed] [Google Scholar]
  • Yang N, Galick H, Wallace SS. Attempted base excision repair of ionizing radiation damage in human lymphoblastoid cells produces lethal and mutagenic double strand breaks. DNA repair. 2004;3(10):1323–1334. [PubMed] [Google Scholar]
  • Ye Y, Stahley MR, Xu J, Friedman JI, Sun Y, McKnight JN, Gray JJ, Bowman GD, Stivers JT. Enzymatic excision of uracil residues in nucleosomes depends on the local DNA structure and dynamics. Biochemistry. 2012;51(30):6028–6038. [PMC free article] [PubMed] [Google Scholar]
  • Yun MH, Hiom K. CtIP-BRCA1 modulates the choice of DNA double-strand-break repair pathway throughout the cell cycle. Nature. 2009;459(7245):460–463. [PMC free article] [PubMed] [Google Scholar]
  • Zhou T, Akopiants K, Mohapatra S, Lin PS, Valerie K, Ramsden DA, Lees-Miller SP, Povirk LF. Tyrosyl-DNA phosphodiesterase and the repair of 3'-phosphoglycolate-terminated DNA double-strand breaks. DNA repair. 2009;8(8):901–911. [PMC free article] [PubMed] [Google Scholar]
-