Sie sind auf Seite 1von 25

Mutation Research 589 (2005) 111–135

www.elsevier.com/locate/reviewsmr
Community address: www.elsevier.com/locate/mutres
Review

DNA–protein crosslinks: their induction, repair,


and biological consequences
Sharon Barker, Michael Weinfeld, David Murray*
Department of Oncology, Division of Experimental Oncology, Cross Cancer Institute, 11560 University Avenue,
University of Alberta, Edmonton, Alberta, Canada T6G 1Z2
Received 10 September 2004; received in revised form 9 November 2004; accepted 17 November 2004
Available online 5 January 2005

Abstract

The covalent crosslinking of proteins to DNA presents a major physical challenge to the DNA metabolic machinery. DNA–
protein crosslinks (DPCs) are induced by a variety of endogenous and exogenous agents (including, paradoxically, agents that
are known to cause cancer as well as agents that are used to treat cancer), and yet they have not received as much attention as
other types of DNA damage. This review summarizes the current state of knowledge of DPCs in terms of their induction,
structures, biological consequences and possible mechanisms of repair. DPCs can be formed through several different
chemistries, which is likely to affect the stability and repair of these lesions, as well as their biological consequences. The
considerable discrepancy in the DPC literature reflects both the varying chemistries of this heterogeneous group of lesions and
the fact that a number of different methods have been used for their analysis. In particular, research in this area has long been
hampered by the inability to chemically define these lesions in intact cells and tissues. However, the emergence of proteomics as
a tool for identifying specific proteins that become crosslinked to DNA has heralded a new era in our ability to study these
lesions. Although there are still many unanswered questions, the identification of specific proteins crosslinked to DNA should
facilitate our understanding of the down-stream effects of these lesions.
# 2004 Elsevier B.V. All rights reserved.

Keywords: DNA–protein crosslink; DNA repair; DNA damage

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

2. Detection of DPCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

3. Chemical-induced DPC formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

* Corresponding author. Tel.: +1 780 432 8430; fax: +1 780 432 8428.
E-mail address: davem@cancerboard.ab.ca (D. Murray).

1383-5742/$ – see front matter # 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.mrrev.2004.11.003
112 S. Barker et al. / Mutation Research 589 (2005) 111–135

3.1. Formaldehyde-induced DPCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114


3.2. Metal-induced DPCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

4. DPCs induced by IR and ROS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115


4.1. Radiation-induced DPCs in cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.2. Radiation-induced DPC structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
4.3. Protein radicals and DPCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

5. Stability of DPCs in vitro . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

6. Biological consequences of DPCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119


6.1. Nickel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
6.2. Chromium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
6.3. Arsenic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
6.4. Formaldehyde . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.5. Methylglyoxal and glyoxal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.6. Pyrrolizidine alkaloids. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6.7. Ionizing radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
6.8. Cumulative/background lesions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

7. Proteins involved in DPCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121


7.1. Crosslinking of DNA replication/repair enzymes to DNA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

8. Crosslinking of DNA to the nuclear matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

9. Enzymatic repair of DPCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125


9.1. How are DPCs sensed at the cellular level? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
9.2. How are covalent DPCs repaired? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
9.3. How are IR-induced DPCs repaired? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
9.4. Might protease activity be involved in DPC repair? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

10. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

1. Introduction various aldehydes, and some important chemother-


apeutic drugs including cisplatin, melphalan, and
The purpose of this review is to summarize our mitomycin C. Humans are continuously exposed to
current understanding of the mechanisms of induction DPC-inducing agents present in environmental pollu-
and repair, as well as the biological consequences, of tants such as cigarette smoke and automotive and
the types of DNA lesion known as DNA–protein diesel exhaust, industrial chemicals and foodstuffs, as
crosslinks (DPCs). A DPC is created when a protein well as physiological metabolites, such as products of
becomes covalently bound to DNA. Such events occur lipid peroxidation. Understanding the biology of these
following exposure of cells to a variety of cytotoxic, lesions is complicated by several factors. For example,
mutagenic, and carcinogenic agents, including ultra- different agents induce DPCs by different mechanisms
violet light and ionizing radiation (IR), metals and (Fig. 1). Proteins can become crosslinked to DNA
metalloids such as chromium, nickel and arsenic, directly through oxidative free radical mechanisms or
S. Barker et al. / Mutation Research 589 (2005) 111–135 113

Fig. 1. Crosslink structures. A schematic representation of two of the chemistries by which proteins may become crosslinked to DNA. (A) A
formaldehyde induced crosslink between cytosine and lysine (taken from [169]). (B) An IR-induced crosslink between thymine and tyrosine
(taken from [44]).

they can be crosslinked indirectly through a chemical 2. Detection of DPCs


or drug linker or through coordination with a metal
atom. A subtype of these crosslinking mechanisms Early studies of DPCs tended to focus on the
involves a sulfhydryl linkage to the amino acid. This issue of whether cellular protein became associated
results in numerous types of DPCs that are chemically with DNA and quantifying these DPCs following
distinct and whose formation is influenced by factors exposure of a test system to a given genotoxic agent.
such as cellular metabolism, cell-cycle phase, and Existing techniques for the quantitation of DPCs
temperature. It is likely that these different types of differ in their detection limit/sensitivity level and
crosslinks will be more or less susceptible to various associated problems. DPC induction can be measured
mechanisms of reversal (e.g., hydrolysis) and enzyme- using the comet assay because the crosslinking of
catalyzed repair, given their different chemical proteins to DNA retards the migration of DNA
structures and physical conformations. They may fragments, resulting in a reduced tail moment [5,6].
also have different cellular consequences. However, this method does not allow for isolation of
The timing of this review coincides with the DPCs. Gradient separation methods (e.g., CsCl,
emergence of proteomics as a tool for studying sucrose) [7,8] separate most DPCs from the bulk
biological complexes involving unknown proteins, of the DNA and protein by density, but DPCs are
so that the identification and quantification of specific found throughout the DNA and protein fractions
proteins that become crosslinked to DNA is now [9].
possible without the necessity for presumption. This A filter-based DPC isolation method employing
approach has been recently highlighted because of its nitrocellulose membranes is useful for obtaining dose
success in identifying proteins involved in complex response curves for total DNA–protein binding based
cellular structures such as the spliceosome [1] and lipid on DNA retention, but is not useful for the
rafts [2]. Such studies have highlighted an important identification of specific proteins involved in DPCs
issue that may have compromised earlier studies of this because nitrocellulose binds all cellular proteins [10–
type, namely that of protein abundance and solubility 12]. A method developed by Zhitkovich and Costa
under a given set of assay conditions, which may [13,14] measures DPC induction as the extent of DNA
greatly influence the proteins that are identified to the associated with protein after the protein is precipitated
exclusion of others. These issues may have contributed using sodium dodecyl sulfate/potassium (SDS/K+).
to discrepancies among earlier studies. However, SDS/K+ precipitation is expected to result in
Two classes of DPC, the attachment of topoisome- the precipitation of some non-covalently linked
rases to DNA and the association of DNA and protein proteins because SDS binds selectively to proteins
caused by hyperthermia, have been reviewed recently and is then precipitated (with bound DNA) by the
[3,4], and will not be discussed in depth in this review. potassium.
114 S. Barker et al. / Mutation Research 589 (2005) 111–135

An alternative approach to DPC quantitation is to this technique does not itself identify the proteins.
isolate DNA and measure the associated protein. The However, the emerging field of proteomics, which
alkaline elution assay traps high molecular weight combines the separating capacity of 2-D PAGE analysis
DNA (with attached proteins) on a polycarbonate filter with powerful protein sequencing technology, should
while non-covalently bound proteins are washed away greatly facilitate the identification of these proteins.
[15,16]. However, recovering DPCs from the filters is
difficult and poorly reproducible. Total genomic DNA
can be isolated using a chaotrope/detergent mix and 3. Chemical-induced DPC formation
ethanol precipitation. This DNA isolation method can
be combined with additional steps to stringently 3.1. Formaldehyde-induced DPCs
dissociate non-covalent protein–DNA complexes to
allow the isolation of proteins truly crosslinked to the Formaldehyde is a widely studied DPC-inducing
DNA. Modifications of this method have been used to agent, and the crosslinking of proteins to DNA by
isolate and identify nuclear matrix proteins cross- formaldehyde is used for the investigation of DNA–
linked to DNA by cisplatin [17,18]. protein interactions in a technique called chromatin
The lack of stringency of DPC isolation methods has immunoprecipitation (ChIP). To perform ChIP, cells are
been part of the problem in assessing the biological treated with formaldehyde resulting in the covalent
relevance of DPC analyses to date. It is known that crosslinking of proteins to the DNA sequences with
nuclear matrix proteins are tightly associated with the which they are associated. The DNA is then fragmented
DNA; their complete dissociation is crucial for the and the protein–DNA complex is isolated by immuno-
identification of those proteins that are covalently precipitation with an antibody to the protein of interest.
crosslinked to DNA by a given agent. As well, proteins Formaldehyde can react with amine, thiol, hydro-
are usually crosslinked at low levels, and it can be xyl, and amide groups to form various types of
difficult to isolate sufficient quantities for the sequen- adducts, but the major class of DNA lesions induced
cing of proteins for identification. Detection limits of by this compound are DPCs (reviewed in [23,24]).
the various techniques have contributed to variability in DPC induction involves the reaction of formaldehyde
results. Several studies have made use of two- with amino and imino groups of proteins (e.g., lysine
dimensional polyacrylamide gel electrophoresis (2-D and arginine side chains) or of nucleic acids (e.g.,
PAGE) to analyze proteins present in crosslinked cytosine) to form a Schiff base, which then reacts with
samples or in nuclear matrix fractions [8,17–22], but another amino group (Fig. 2) [25,26].

Fig. 2. Formaldehyde crosslinking mechanism. This figure depicts the steps in the reaction of formaldehyde with an amino group (e.g., of a
protein side chain) to form a Schiff base (in step 1) which can then go on and react with another amino group (of a DNA base) to complete the
crosslink.
S. Barker et al. / Mutation Research 589 (2005) 111–135 115

3.2. Metal-induced DPCs separated from unreacted Cr(III) and reacted with
protein. Additionally, these investigators reacted the
Among the DPC-inducing agents commonly found Cr(III)–histidine complex with nucleosides and nucleo-
as environmental and workplace pollutants are a tide monophosphates and showed that nucleotides
number of metal compounds. DPCs induced by nickel could participate in crosslinks but nucleosides could
compounds have been suggested to involve oxidative not, indicating that the phosphate group is essential for
mechanisms [27,28]. Nickel ions have a high affinity for the crosslinking reaction. However, this crosslinking
proteins, especially for histidine, cysteine, and aspartic utilized free amino acid and free nucleotide and thus
acid residues [27,28]. In one study [27], DPCs were may not be identical to that which would occur in vivo.
isolated by SDS/K+ precipitation from rat lymphocytes The different types of linkages seen with chromium
treated with various nickel compounds. Co-incubation treatment (chelation complexes, sulfhydryl linkages,
of lymphocytes with nickel compounds and either metal and linkages generated by ROS) raise the interesting
chelators, free amino acids, or scavengers of reactive question as to whether other DPC-inducers can generate
oxygen species (ROS) all decreased the yield of DPCs. more than one type of crosslink and what factors might
Analysis of metal ion-induced crosslinks demon- influence the spectrum and yield of various types of
strated that not all putative DPCs are due to covalent crosslinks produced by a given agent.
linkages [22] and that one agent can induce more than
one chemical type of crosslink. DPCs were induced in
human leukemic cells or isolated nuclei by treatment 4. DPCs induced by IR and ROS
with potassium chromate, chromium (III) chloride or
IR. DPCs were isolated by SDS/K+ precipitation/ 4.1. Radiation-induced DPCs in cells
ethanol precipitation and analyzed by 2-D SDS-
PAGE. Some crosslinked proteins were liberated by Exposure of cells to IR results in the generation of
treatment with EDTA, indicating that they were not many localized ROS within a short distance of each
covalently crosslinked to DNA but rather were bound other and of the DNA (Fig. 3). Many of these,
to DNA through a chelatable form of chromium. Some including the extremely reactive hydroxyl radical
crosslinked proteins were liberated by treatment with (OH), will be generated at high levels within small
thiourea, indicating that they were crosslinked to DNA discrete regions known as spurs, blobs, and short
through a sulfhydryl linkage. The majority of IR- tracks [30]. When these ionization-dense regions
induced DPCs were not reversed by EDTA or thiourea overlap a DNA molecule, this can result in what are
treatment and were only released from the DNA by
DNase I digestion, and likely represent covalent
crosslinks formed through oxidative mechanisms.
Some of the DPCs induced by chromate were also
resistant to EDTA or thiourea treatment, and were thus
likely to be covalent linkages formed via ROS.
Zhitkovich et al. [29] reported that a considerable
proportion (50% at biologically relevant doses) of
chromium–DNA adducts were in fact DNA–metal–
protein complexes. The amino acids most frequently
involved in these complexes were cysteine, histidine,
and glutamic acid. Reactions of cysteine or histidine
with trivalent or hexavalent chromium were analyzed, Fig. 3. Generation of ROS by IR: IR can directly ionize DNA or
and it was shown that Cr(VI) must be reduced to Cr(III) protein in its path generating DNA or protein radicals. Indirectly
and that Cr(III) must first complex with an amino acid ionizing events include the ionization of water molecules surround-
ing the DNA or protein generating the reactive hydroxyl radical
before reacting with DNA to form the crosslink. No (OH), which can then react with DNA or protein, rendering it
complex was formed between DNA and amino acid if reactive. The dashed circle represents a spur (Section 4.1) [30]. The
the DNA alone was first incubated with Cr(III) and then shaded globules represent proteins.
116 S. Barker et al. / Mutation Research 589 (2005) 111–135

variously referred to as ‘‘locally multiply damaged


sites’’ or ‘‘clustered lesions’’, because each radical
within the region can potentially generate damage to
the DNA. The result is multiple types of damage—
single strand breaks (SSBs), double strand breaks
(DSBs), base damage or base loss, DNA–DNA
crosslinks, and/or DPCs—generated within a short
distance of each other in the DNA. Most studies of the
biological effects of the cellular lesions induced by IR
have focused on DSBs, and not much attention has
been paid to the DPC. However, measurements of the Fig. 4. Oxygen dependence of DPCs (squares) and survival (trian-
amounts of each type of damage induced per gles) in g-irradiated AA8 CHO cells. DNA–protein crosslinks were
mammalian cell per unit absorbed dose of IR reveal measured by the alkaline elution assay and cell survival was
measured by colony-forming assay. The x-axis represents the per-
that the yield of DPCs (150/cell per Gy) is actually cent oxygen in the gassing mixture. Single-cell suspensions were
higher than that of either DSBs (20–40/cell per Gy) or stirred at 4 8C while being gassed with a mixture of 5% CO2, varying
DNA–DNA crosslinks (30/cell per Gy) [31]. concentrations of O2, balance N2, for 3 h prior to irradiation.
Early studies by Fornace and Little [32,33] using
alkaline elution demonstrated the induction of DPCs in
aerated human cells exposed to very high doses of X- Given that the yield of DPCs in cells decreases
rays. They also showed an increase in DPC induction markedly as oxygen is introduced, whereas the effect
efficiency under hypoxic conditions. A similar obser- of oxygen on IR-induced cell killing goes in the
vation was made by Meyn et al. [34,35] using Chinese opposite direction, and because the yield of other types
hamster ovary (CHO) cells and by Radford [36] using of DNA damage such as DSBs closely parallels cell
mouse L cells, again using alkaline elution, and by Xue killing under these conditions, the role of DPCs in the
et al. [11] in V79 hamster cells using a filter binding biological effects of IR has been largely disregarded.
assay. Zhang et al. [37–39] suggested that negligible However, as will be discussed in Section 9, these
levels of DPCs are formed at oxygen concentrations lesions may contribute to the radiosensitivity of
above 1%, that there is maximal DPC induction at oxy- hypoxic cells if their repair is compromised.
gen concentrations below 0.1%, and that oxygenated The situation with respect to DPCs and high linear
cells are 10–100-fold less susceptible to forming DPCs energy transfer (LET) radiation has received some
than hypoxic cells. Similarly, vanAnkeren, Murray, and theoretical consideration. One unresolved question is
Meyn (unpublished data) examined the relationship whether DPCs, either alone or in association with
between oxygenation and DPC induction in CHO cells clustered lesions, might differentially contribute to
exposed to g-radiation and found that the yield of DPCs cell killing induced by radiations of differing LET.
decreases as oxygen levels increase (Fig. 4). Several The thinking is that higher LET tracks will generate
other studies have also shown a marked increase in more complex clustered lesions, possibly with a
cellular DPCs induced by IR under hypoxic conditions higher probability of involving a DPC. Putative high-
[40–44]. LET ‘‘specific’’ lesions could include complex
Zhang et al. [38] showed that pH, nutrient depletion, clustered-damaged sites wherein DSBs are associated
temperature, and growth phase did not significantly with DPCs [46–48]. Some experimental studies have
influence the yield of IR-induced DPCs in aerated addressed the issue of whether the yields and/or repair
normal and tumor cells as measured by alkaline elution. of DPCs might differ with LET. Blakely et al. [49]
Similarly, pH and nutrient status had no effect on showed that the initial DPC yields in normal hamster
cellular DPC induction when oxygen was absent [45]. cells were similar for X-rays and high energy Ne-ions
Importantly, Zhang et al. [38] pointed out that it is of 32, 100, and 183 keV/mm at low doses, although N-
difficult to compare DPC studies as the various ions (120 keV/mm) generated a lower DPC yield.
techniques used to measure DPCs differ in their Another study suggests that a high-LET beam of N-
detection limits. ions appeared to induce higher levels of residual (6 h
S. Barker et al. / Mutation Research 589 (2005) 111–135 117

post-IR) DPCs per unit dose than low-LET X-rays in particular drugs and environmental pollution. The
human melanoma cells (data from Eguchi et al. [50], crosslinking mechanism involves H-atom abstraction
re-calculated by Frankenberg-Schwager [51]). This from the methyl group of thymine by OH, addition of
difference may be attributable to the above-mentioned the resultant thymine radical to the carbon-3 position
induction of more complex lesions at higher LET, of the tyrosine ring, and oxidation of the resulting
rendering DPCs more difficult to repair. adduct radical [54].
Electrospray–ionization mass spectrometry (ESI–
4.2. Radiation-induced DPC structures MS) analysis of an irradiated solution containing
angiotensin and thymine demonstrated the formation
To understand the cellular consequences of DPCs of a covalent bond between the methyl group of
and to investigate their possible repair pathways, it thymine and C3 of the angiotensin tyrosine ring [57]
will be important to delineate the chemistries of these and also indicated C2 of tyrosine as another major site
linkages (Fig. 1). Extensive work with cell-free of bond formation. Crosslinks between thymine and
models has demonstrated the covalent nature of IR- tyrosine were detected at IR doses as low as 0.1 Gy,
induced DPCs, and the chemical structure of some and the yield of crosslinks was linear up to 100 Gy.
DPCs has been determined using gas chromatography/ Reaction of OH with thymine most frequently
mass spectrometry (GC/MS) analyses [52–54]. These resulted in addition to the C5–C6 double bond
reports examined g-irradiated aqueous mixtures of (60% and 30%, respectively, at the 5 and 6
thymine and amino acids (lysine, glycine, alanine, positions), and abstraction of an H-atom from the
valine, leucine, isoleucine, tyrosine, and threonine) methyl group occurred only 10% of the time.
and demonstrated that particular DNA–amino acid It will be of interest to determine if specific proteins
crosslinks exist as several isomers [52–54]. The found to be covalently crosslinked to DNA in vivo will
involvement of these amino acids in DPCs was also prove to be linked through any of these identified
shown in vitro in isolated irradiated mixtures of calf target residues. Additionally, this information may be
thymus nucleohistone [52–54]. of use in predicting which proteins are likely targets
The GC/MS experiments were extended to analyze for DPC formation because of their amino acid
the formation of DPCs in vivo using cultured composition and their contact with the DNA.
mammalian cells [44,55] and rat renal tissue [56]. Identifying a crosslinked protein and the residue
These samples were treated with ferrous ions, through which the linkage forms may also provide
hydrogen peroxide, or IR, and the chromatin was information on molecular geometry because the DNA
isolated, subjected to acid hydrolysis, and analyzed by and protein must be in close proximity during free
GC/MS. Crosslinking of DNA to protein through a radical generation.
thymine–tyrosine linkage was detected in these
samples. In both the in vitro and in vivo studies, the 4.3. Protein radicals and DPCs
induction of DNA–amino acid complexes and DPCs
increased linearly with IR dose. Hydrogen peroxide DNA is not the only site of free radical generation
treatment of cultured cells also resulted in the or the only target for free radical attack following IR
concentration-dependent induction of DPCs in chro- exposure (Fig. 3). Proteins and amino acids are also
matin [44]. Addition of radical scavengers/metal susceptible to attack by ROS. Indeed, an alternative
chelators (dimethylsulfoxide or o-phenanthroline) mechanism for DPC induction involves an initial
partially inhibited DPC formation [44]. protein radical created by abstraction of an H-atom by
Dizdaroglu et al. [54] have proposed that the OH 
OH from the amino acid, followed by addition of the
radical is involved in the formation of the crosslink amino acid radical to the C6 position of thymine and
whether these DPCs are induced by ferrous ions, oxidation of the adduct radical [52]. ESI–MS studies
hydrogen peroxide or IR. Free radicals/ROS are also by Weir Lipton et al. [57] show that OH adds to the
generated through biological redox reactions and tyrosine ring at C3 50% of the time and at C2 35%
under conditions causing oxidative stress, such as of the time. The C3 tyrosine adduct radical then loses
malnutrition, numerous disease states, exposure to water to generate a phenoxyl radical, which can then
118 S. Barker et al. / Mutation Research 589 (2005) 111–135

react with DNA. Thus, a DPC may be formed by the formation was decreased in the presence of radical
addition of a protein radical to DNA or vice versa, or scavengers suggested that a radical is involved in this
from a combination of two radicals. reaction.
Exposure of proteins to ROS can generate protein Similar steps in DPC formation were suggested by
hydroperoxides or other reactive protein species as analysis of malondialdehyde-induced DPCs in vitro
well as additional free radicals. An in vitro study [58] [59]. These investigators reacted malondialdehyde
used several purified proteins (insulin, a-casein, with either protein or DNA in aqueous solution,
apotransferrin, and bovine serum albumin (BSA)) purified away non-reacted material, and then
irradiated in aqueous solution in the presence of attempted the second half of the DPC reaction (by
oxygen or nitrous oxide to generate protein hydro- introducing DNA or protein). For the formation of a
peroxides, and tested these for DPC formation with DPC, it was apparent that the malondialdehyde must
plasmid DNA based on the retardation of DNA first react with the protein to generate an adduct that
migration on an agarose gel. The observation that subsequently reacts with the DNA to form the
inclusion of anti-oxidants did not reduce the yield of crosslink.
DPCs suggested that these lesions were not generated Although both DNA- and protein-radical formation
from long-lived radical species produced at the have been suggested as the first step in DPC formation
irradiation step. However, the formation of DPCs in vitro, it remains to be seen which mechanism
was reduced by including metal chelators in the predominates in vivo. Both mechanisms are probably
reaction, suggesting that at least some of the DPCs operative for various agents, and other factors may
were dependent on metal atoms associated with the influence the levels of each type of radical produced.
DNA. Other reports have indicated that proteins that For example, in the case of IR, the spatial distribution
do not bind to DNA (e.g., BSA [59]) cannot generate of DNA and proteins in the radiation track may be
DPCs in vitro, so there is some question as to whether critical in this regard [48].
or not non-DNA binding proteins can be involved in
DPCs. It is likely that the conflicting reports reflect
variations in in vitro experimental parameters such as 5. Stability of DPCs in vitro
DNA and/or protein concentrations, presence of
radical scavengers, and presence of salts or metals Different types of DPCs appear to have very
or reductants that would interfere with the DPC different chemical stability. Aldehyde-induced DPCs
formation reaction. are reversed by spontaneous hydrolysis and are also
Further work examining the role of reactive reversible by incubation at elevated temperatures
protein species in DPC formation used hypochlorous (discussed in [24]). Acetaldehyde-induced DPCs are
acid (HOCl), an oxidant that is produced by normal hydrolytically unstable, and in in vitro experiments
metabolic processes such as phagocyte activity [60]. only 25% of these DPCs remained after 8 h at 37 8C
HOCl can react with protein amino groups, generat- [61,62]. By comparison, malondialdehyde-induced
ing chloramines that decompose to protein radicals, DPCs formed in vitro using purified DNA and histone
which can react with DNA. HOCl can also interact protein had a much longer half-life of 13.4 days at
with DNA to form chloramines. Hawkins et al. [60] 37 8C [59].
investigated the formation of DPCs by HOCl in The lifetime of formaldehyde-induced DPCs in
nucleosomes of eukaryotic-cell nuclei using electron vitro was investigated by Quievryn and Zhitkovich
paramagnetic spin resonance spectroscopy. The [24] using purified DNA and histone H1. Inclusion of
reaction of protein radicals with pyrimidine nucleo- either SDS or 0.8 M sodium chloride with the
sides was observed to yield nucleobase radicals, formaldehyde during the crosslinking reaction
which could result in covalent crosslinking of DNA to reduced crosslinking of histone H1 by preventing its
protein. These authors [60] suggested that reaction of binding to DNA. Addition of SDS after the
HOCl occurs predominantly with the protein and not formaldehyde crosslinking reaction decreased the
the DNA, and that 50–80% of these reactions are with lifetime of the histone H1–DNA DPC from 26.3 to
lysine or histidine residues. The finding that adduct 18.3 h at 37 8C, suggesting that if the protein is
S. Barker et al. / Mutation Research 589 (2005) 111–135 119

allowed to stay associated with the DNA, the 6.2. Chromium


crosslinks can reform under physiological conditions.
Chromium exposure has been associated with an
increased incidence of respiratory cancers (reviewed
6. Biological consequences of DPCs in [74]). Voitkun et al. [75] used amino acid–
chromium–DNA adducts (model DPCs) in a shuttle
The covalent crosslinking of proteins to DNA is vector to show that processing of these lesions by
expected to interrupt DNA metabolic processes such human cells can result in mutagenesis. Plasmids
as replication, repair, recombination, transcription, containing DNA–Cr(III)–glutathione or DNA–
chromatin remodeling, etc. Indeed, the effect of agents Cr(III)–amino acid adducts were transfected into
that cause DPCs on DNA replication has been widely human fibroblasts, re-isolated after a 48 h incubation,
investigated ([63–65] and others). DPCs are expected and sequenced. The types of mutations caused by the
to act as bulky helix-distorting adducts and would DPCs were mainly single base substitutions at G:C
therefore be likely to physically block the progression base pairs, with G:C ! A:T transitions and
of replication or transcription complexes and/or G:C ! T:A transversions being induced with similar
prevent access of proteins required either for synthesis frequency. Chromium–DNA complexes also resulted
along the template strand, for transcription, or for in sequence mutations, although this effect was
repair recognition and/or incision. They may also weaker.
affect all of these processes by anchoring the The feasibility of using DPCs as biomarkers for
chromatin and preventing its remodeling. exposure to chromium in human cells has been
Unfortunately, our understanding of the biolo- investigated [74]. Higher levels of DPCs were
gical consequences of DPCs is hampered by the fact detected in lymphocytes of individuals exposed to
that no agent exclusively induces these lesions in chromium compounds than in non-exposed indivi-
genomic DNA (although studies using plasmid DNA duals, although the DPC level was found to plateau in
have provided some insight into the processing of individuals exposed to high levels of chromium.
these lesions by cells; see below). Thus, all known
DPC-inducing agents generate other forms of DNA 6.3. Arsenic
damage in addition to DPCs, and direct attribution of
any observed effect such as mutagenesis or carcino- Arsenic has been implicated in the induction of
genesis to DPCs is inevitably confounded by the skin, lung, bladder, and liver cancers [76–78].
concomitant impact of these other lesions. None- Although it is carcinogenic, arsenic has not been
theless, several studies have reported that the induc- found to be mutagenic. Earlier studies suggested that
tion of DPCs by many agents correlates with genetic arsenic only induces DNA damage at high concentra-
damage such as sister chromatid exchanges (SCEs), tions; however, a recent study [79] suggests that
transformation, and cytotoxicity [66–70]. Thus, DPCs different cell types differ in their sensitivity to arsenic.
may contribute to the genotoxic effects of many Arsenic does in fact induce DNA damage at
different DNA-damaging agents, some of which are concentrations that are biologically relevant, the
discussed below. major forms of arsenic-induced DNA damage being
oxidative DNA adducts and DPCs [79]. As well,
6.1. Nickel multiple pathways have been proposed for arsenic-
induced cytotoxicity [79]. Treatment with arsenite
Various types of chromosome damage (DNA gaps may result in DNA damage through the production of
and breaks, SCEs and others) have been shown to HOCl because there is an activation of NADH oxidase
persist in lymphocytes of nickel workers for years and an increase in superoxide production after NADH
after exposure [71,72]. Earlier studies demonstrated addition in arsenite-treated human vascular smooth
an increased incidence of alveolar/bronchial/adrenal muscle cells [80]. This pathway can result in DNA
medulla neoplasms in rats exposed to nickel com- damage because superoxide is converted to hydrogen
pounds [73]. peroxide by superoxide dismutase, and the resulting
120 S. Barker et al. / Mutation Research 589 (2005) 111–135

hydrogen peroxide can react with chloride ions to and may have similar applicability in assessing risk
form HOCl or with transition metal ions to produce factors for exposure to other DPC-inducers.

OH [80–82].
Evidence that arsenic cytotoxicity may not be due 6.5. Methylglyoxal and glyoxal
to DNA damage comes from Mei et al. [83]. Similar
sensitivity was seen for normal human cells and Methylglyoxal [pyruvic aldehyde: CH3COCHO] is
various DNA repair-deficient cell lines (Xeroderma another endogenous aldehyde metabolite known to
Pigmentosum (XP), Bloom Syndrome (BS), and induce DPCs. It is found widely in food and beverages
Fanconi Anemia (FA)) after treatment with sodium and in cigarette smoke. Methylglyoxal reacts with free
arsenite; however, Ataxia-Telangiectasia (AT) cells amino acids, proteins, and nucleic acids (mainly
were significantly more sensitive. This sensitivity did guanines) resulting in DNA adducts, strand breaks,
not appear to be related to DSB repair because DNA interstrand crosslinks, and extensive DNA–pro-
additional cell lines defective in DSB repair did not tein crosslinking through lysine and cysteine residues
display increased sensitivity to arsenic. As well, there [91], including crosslinking of histones (reviewed in
was no induction of DSBs (as measured by histone [92–94]). Mutations induced by methylglyoxal in
H2AX phosphorylation) and no activation of p53 upon mammalian cells were predominantly (50%) dele-
treatment of normal cells with sodium arsenite. One tions but included a significant proportion of base-pair
parameter that did seem to be affected by arsenic substitutions (35%) [93]. The DNA-damaging effects
treatment was cell cycle distribution. Normal cells of methylglyoxal include the induction of SCEs,
showed a significant increase in the percentage of cells chromosomal aberrations, and micronuclei [93].
in S-phase and a modest increase in the percentage of Glyoxal [(CHO)2] is a related, endogenously
cells in G2/M phase after arsenic treatment, whereas produced, aldehyde that induces DNA strand breaks
the cell cycle distribution of AT cells was unaffected. but 10-fold fewer DPCs than methylglyoxal.
Thus, the sensitivity of AT cells to arsenic may be due Glyoxal also induces 10-fold fewer frameshift
to an effect on cell cycle regulation and not necessarily mutations than methylglyoxal, suggesting that DPCs
due to DNA damage. However, Bau et al. [79] might be the cause of these events (which are a
provided evidence that arsenic induces DPCs that are common result of bulky adducts) [94]. Roberts et al.
converted to DSBs over time. Thus, measurements of [94] compared the effects of glyoxal and methyl-
DSBs and DPCs will be inaccurate as DPCs become glyoxal on human skin cells using both the comet
converted to DSBs. The disruption of cell cycle seen assay and an in vitro plasmid assay. In the comet assay,
with arsenic treatment may be due to DPCs. Although the tail moment increased when cells were treated with
there is little knowledge on the effect of DPCs on cell glyoxal, indicating DNA strand breakage. However,
cycle progression, these lesions are expected to disrupt following methylglyoxal treatment, there was com-
multiple functions of DNA metabolism/organization. paction of the nucleus and reduced migration,
indicating the presence of DPCs.
6.4. Formaldehyde
6.6. Pyrrolizidine alkaloids
Formaldehyde [HCHO] is the most widely studied
DPC-inducing agent. It is mutagenic in bacteria, lower Pyrrolizidine alkaloids are cytotoxic compounds
eukaryotes, and human lymphoblasts, inducing pri- found in many plant species that are used in herbal
marily point mutations and deletions. Formaldehyde remedies and teas. These compounds can cause liver
also causes micronuclei [84] and is implicated in the disease and are carcinogenic [95]. They are metabo-
induction of nasal tumors in experimental animals lically activated and form DPCs and DNA interstrand
[85,86]. The induction of DPCs by formaldehyde has crosslinks in similar proportions when assessed by
been shown to be dose-dependent and to correlate alkaline elution [96]. The cytotoxic and anti-mitotic
with tumorigenesis [87,88]. The extent of DNA– activities of pyrrolizidine alkaloids correlates with
protein crosslinking has been used as a biomarker of their ability to form both DPCs and interstrand
formaldehyde exposure in mammalian cells [87,89,90] crosslinks [96–98].
S. Barker et al. / Mutation Research 589 (2005) 111–135 121

6.7. Ionizing radiation are bound may help to unravel the biological
consequences of DPCs as well as the mechanisms
As noted in Section 4.1, the role of DPCs in the of their repair. A number of investigators have tried to
biological effects of IR has been largely ignored identify proteins that can become crosslinked to DNA
because these lesions are more abundant following using in vitro systems with purified proteins and DNA
irradiation in the absence of oxygen, a condition that is or by isolating DPCs from cells exposed to various
protective for most other IR-induced end-points such DNA-damaging agents. Several proteins have been
as cell killing and mutation. Certainly, this observation shown to be amenable to crosslinking in vitro when
suggests that DPCs are minor lesions in irradiated they are combined with DNA and treated with a DPC-
oxygenated cells. However, there is some evidence inducing agent, although the relevance of this infor-
that DPCs can contribute to the killing of mammalian mation to the in vivo situation is uncertain. Some
cells when their repair is inhibited. In particular, reports suggest that only DNA-binding proteins can be
certain DNA repair-deficient hamster cell lines such as crosslinked to DNA, while others suggest that any
UV41 (XPF) and UV20 (ERCC1) (reviewed in protein can become crosslinked to DNA. Potentially
[99,100]) are significantly more sensitive than wild- biologically-relevant proteins that have been shown to
type cells to killing by IR under hypoxic conditions, a be crosslinked to DNA in vivo include actin,
phenotype that has been attributed to a deficiency in lectin, aminoglycoside nucleotidyl transferase, his-
the repair of DPCs [100]. tones, a heat shock protein (GRP78), cytokeratins,
It should be noted that many human tumors contain a vimentin, protein disulfide isomerase, and transcrip-
significant proportion of hypoxic cells, and this tion factors/co-factors (estrogen receptor, histone
represents a problem in the use of radiation therapy deacetylase 1, hnRNP K, HET/SAF-B) (Table 1)
for cancer treatment because hypoxic cells are more [7,18,19,22,102–105].
resistant to IR-induced killing. The findings that DPCs Actin was shown to be crosslinked to DNA in
are induced by IR to a greater extent in hypoxic versus human leukemic cells or isolated nuclei treated with
aerated cells and that certain repair deficiencies specifi-
cally increase the radiosensitivity of hypoxic cells Table 1
might provide an avenue for improving radiation Proteins identified in DNA–protein crosslinks
therapy if the repair of DPCs can be effectively Protein Crosslinking agent Reference
inhibited. Actin Chromium [19,22]
Cisplatin [106]
6.8. Cumulative/background lesions Mitomycin C [106]
Pyrrolizidine [106]
Alkaloids
DPC accumulation may be associated with breast
Lectin Chromium [22]
cancer [101]. The base-level of DPCs, presumably Aminoglycoside Chromium [22]
caused by environmental factors and metabolic nucleotidyl transferase
byproducts, was found to be significantly elevated Histones H1, H2A, Formaldehyde [104]
in breast cancer patients compared to healthy H2B, H4
Histone H3 Formaldehyde [104]
individuals. It is far from clear, however, whether
Gilvocarcin V [102]
these DPCs are secondary to the many cellular Glucose regulated Gilvocarcin V [102]
changes that accompany cancer development or protein 78
treatment or if these DPCs are in fact causative in Cytokeratins Arsenic [103]
breast carcinogenesis. Vimentin Formaldehyde [7]
Metabolic byproducts [7]
Protein disulfide Cisplatin [105]
isomerase
7. Proteins involved in DPCs Estrogen receptor Cisplatin [18]
HET/SAF-B
Determining which proteins become crosslinked to hnRNP K
Histone deacetylase 1
DNA by these various genotoxic agents and how they
122 S. Barker et al. / Mutation Research 589 (2005) 111–135

chromium compounds or IR [21,22]. DPCs were One protein identified as being closely associated
isolated by SDS/K+-urea precipitation/ethanol pre- with DNA in vivo by virtue of its susceptibility to
cipitation, followed by analysis by 2-D SDS-PAGE. In crosslinking by formaldehyde is vimentin, which is a
this study, 20 proteins were found to be crosslinked structural/scaffold protein [7]. DPCs were isolated
to DNA by chromium and IR. Three of these were from formaldehyde-treated mouse and human cells by
identified as actin, aminoglycoside nucleotidyl trans- sucrose gradient sedimentation followed by repeated
ferase, and lectin. Similarly, Miller et al. [19] SDS/K+ precipitation/ethanol precipitation, followed
demonstrated the crosslinking of actin to DNA in by immunoprecipitation using anti-vimentin antibo-
hamster cells exposed to chromium or cisplatin. DPCs dies. The vimentin could be released from the DPC by
were isolated by SDS/K+-urea precipitation/acetone boiling, which may indicate thermolability of the
precipitation. DNA was digested with DNase I, and the crosslinkage or a non-covalent association. Vimentin
isolated proteins were analyzed by SDS-PAGE. This DPCs were also observed in oxidatively-stressed and
procedure isolated several proteins, one of which was senescent cells, indicating that metabolic byproducts
identified as actin on the basis of molecular weight and can crosslink this protein to DNA.
pI, and confirmed using immunological methods. Gilvocarcins are naturally occurring anti-tumor
Actin–DNA crosslinks comprised 20% of the total antibiotics that can crosslink proteins to DNA. Normal
DPCs isolated. Additional proteins were found to be human fibroblasts treated with gilvocarcins were
crosslinked by chromium at higher metal concentra- subjected to lysis and DPC isolation using SDS/K+
tions. precipitation with a sodium chloride wash step,
Actin was also found to be crosslinked to DNA by followed by immunoprecipitation with an antibody
pyrrolizidine alkaloids [106]. Bovine kidney cells and to double stranded DNA [102]. The DPCs were
human breast cancer cells were treated with these separated by SDS-PAGE, and two proteins—histone
compounds, and DPCs were isolated by repeated H3 and heat shock protein GRP78—were identified by
extraction/precipitation with SDS and urea. Cross- amino-terminal amino acid sequencing and confirmed
linked proteins were released from the DNA by by immunoblotting [102].
DNase I digestion and analyzed by SDS-PAGE. There are conflicting reports regarding the involve-
Participation of different isoforms of actin in DPCs ment of histones in DPCs. Several investigations have
was confirmed by immunoblotting. Actin was also focused on the in vitro induction of histone-involving
identified as a component of DPCs isolated from DPCs in aqueous solution. Miller et al. [19] treated a
cells treated with cisplatin or mitomycin C. Another combination of purified actin or histone and bacter-
study [105] demonstrated the cisplatin-induced iophage DNA with chromium compounds in vitro and
crosslinking of at least four proteins to DNA in found that histones were not as efficiently crosslinked
human cells and identified protein disulfide isomerase to DNA as actin. This may be due to the fact that
as one of these using immunological methods. If the chromate has a high affinity for sulfhydryl groups and
association of proteins with DNA was disrupted by thus induces crosslinks through a sulfhydryl linkage,
extracting the cells with dithiothreitol prior to but there are few sulfhydryl groups in histone proteins
cisplatin treatment, protein disulfide isomerase was [21]. However, histones have been found to be readily
no longer crosslinked. Several proteins have been crosslinkable to DNA by formaldehyde through an
shown to be crosslinked to DNA by arsenic [103]. amine to amine linkage [104,107,108] and mamma-
DPCs were isolated from arsenic-treated cultured lian histones can be crosslinked to DNA by treatment
human hepatic cells using SDS/K+ precipitation with aldehydes both in treated cells and in cell-free
(without urea). Crosslinked proteins were separated systems [24,59,61,109,110]. The choice of DPC-
by SDS-PAGE, and the presence of several different inducing agent may explain why some studies found
cytokeratins was confirmed using antibodies. How- histones to be highly crosslinked to DNA while others
ever, these arsenic concentration-dependent cross- did not.
links could be reversed by high salt, suggesting that Induction of DNA–histone crosslinks by IR has
they may be non-covalent associations rather than proven controversial. Several studies [52–54] have
true covalent DPCs. shown the IR dose-dependent crosslinking of histones
S. Barker et al. / Mutation Research 589 (2005) 111–135 123

to DNA in vitro using calf nucleohistone. Studies from tumor drug (FR-66979) covalently crosslinks a DNA
Xue et al. [111] and Oleinick et al. [112] using duplex with a synthetic peptide corresponding to the
irradiated hamster cell nuclei demonstrated that DPCs HMGA (formerly HMGI/Y [120]) binding domain.
were induced in histone-depleted chromatin [112] and Extending this work, Beckerbauer et al. [121] reported
that extraction of nuclei with 1.6 M NaCl showed little the crosslinking of HMGA and of HMGB1 and
depletion of DNA-associated histones but was HMGB2 (formerly HMG1 and HMG2 [120]) to DNA
associated with a significant decrease in DPC in vivo by a related drug (FR900482). Complexes of
induction, indicating that other proteins are involved HMGA and DNA were isolated from drug-treated
in these DPCs [111]. However, Mee and Adelstein cells but not control cells using a modified ChIP
[41] also examined the induction of DPCs by g- procedure and HMGA antibodies. In this study, the
radiation using chromatin isolated from Chinese ‘‘crosslinked’’ protein was released from the DNA by
hamster lung fibroblasts and obtained different results. proteinase K digestion, making it difficult to
They suggested that the core histones (H2A, H2B, H3, determine if these complexes were in fact covalent.
and H4) are in fact the major proteins involved in Although the affinity of HMGB1 for undamaged DNA
DPCs because they observed no difference in is very weak, it does have very high affinity for
induction of DPCs between in vitro-prepared whole unusual DNA structures [113]. HMG proteins bind
chromatin and chromatin stripped of other nuclear tightly to chromium-damaged DNA and HMG–Cr–
matrix proteins. These contradictory results may be DNA complexes are stable in 0.5 M NaCl [122], and
due to differences in the efficiencies of the extraction the affinity of HMGB2 for cisplatin-modified DNA is
procedures, and thus the true extent of the involvement 10-fold stronger [123].
of histones in DPCs is yet to be resolved. The question of whether or not HMG proteins are
The conflicting data on the formation of histone– involved in DPCs requires further investigation. HMG
DNA crosslinks may reflect the fact that these studies proteins are known to be extremely mobile [113–116]
used different methods of inducing, isolating, and and, although they are highly abundant and frequently
quantitating DPCs. Given that DPC-inducing agents associated with DNA, their association with DNA
have different mechanisms of action, it is possible that could be too transient for them to be ‘‘trapped’’ in the
histones are substrates for only some types of crosslinking reaction. The above-mentioned affinity of
reactions. Different methods of isolation and analysis these proteins for damaged DNA may favor such
may result in a failure to detect crosslinked proteins of reactions during extended treatments, increasing the
low abundance, and detectability may be affected by likelihood of a crosslinking event.
the solubilities of these proteins. These types of
problems are also likely to affect the analyses of other 7.1. Crosslinking of DNA replication/repair
proteins involved in DPCs. enzymes to DNA
Like the histone proteins, high mobility group
(HMG) proteins are likely targets for DPC induction The potential for crosslink formation between
given that they are highly abundant and frequently DNA replication/repair proteins and the substrate
associated with DNA. These proteins have roles in DNA has been demonstrated by in vitro experiments.
modifying the compaction of the chromatin fiber, HOCl is capable of crosslinking purified DNA single-
promoting access to nucleosomes, and stimulating stranded binding protein to single-stranded oligonu-
transcription and replication [113–116]. Additionally, cleotides in vitro [124]. Methylglyoxal was similarly
the high affinity of HMG proteins for unusual shown to crosslink purified Klenow fragment to a
structures (e.g., chromium-damaged, cisplatin- synthetic DNA substrate [93]. The 2-deoxyribono-
damaged DNA) may also predispose them to cross- lactone lesion is an abasic site produced by a variety of
linking. There is little evidence for the involvement of DNA damaging agents, including IR. This lesion and
HMG proteins in DPCs. HMG proteins were shown to its b-elimination product were prepared in a synthetic
be crosslinked in vitro to a synthetic nitric oxide- substrate and incubated in separate reactions with
damaged DNA substrate [117] (as discussed in Section protein (Escherichia coli endonuclease III, endonu-
7.1). It has been shown [118,119] that a novel anti- clease VIII, FPG (formamidopyrimidine glycosylase),
124 S. Barker et al. / Mutation Research 589 (2005) 111–135

or NEIL1 (a mammalian DNA glycosylase [125])) 8. Crosslinking of DNA to the nuclear matrix
resulting in the crosslinking of each of these proteins
to the lesions [126]. Another study demonstrated that The nuclear matrix is a three-dimensional network
the 2-deoxyribonolactone lesion could be crosslinked that is necessary for DNA organization and nuclear
to DNA polymerase b [127]. structure and function. This framework consists of the
Nitric oxide (NO) is a product of inflammation, and nuclear membrane with the nuclear lamina and pore
chronic inflammation is a known risk factor for many proteins, the internal network of ribonuclear proteins,
cancers. NO-induced damage includes DPCs [128– and nucleolar proteins [134]. The nuclear matrix
130]. One type of DNA damage induced by the contains anchoring sites for the DNA called matrix
nitrosation of guanine by NO is oxanine (Oxa). A attachment regions (MARs) and the DNA is organized
synthetic duplex DNA containing Oxa was shown to into loops of 50–200 kbp between these anchor sites.
form covalent crosslinks between the Oxa moiety and Loop domain anchoring allows for differential control
DNA repair proteins [117]. The E. coli DNA repair of supercoiling between loops during processes such
proteins endo VIII, Fpg, AlkA, and mammalian as replication and transcription [135] which are known
hOGG1 (which bind such types of base damage) to alter DNA topology.
formed DPCs rapidly, while histones and HMG Nuclear matrix proteins are associated with
proteins formed DPCs more slowly and the E. coli processes such as DNA replication, transcription,
Endo III and mammalian hNTH1 and mMPG did not and repair [134]. Some proteins isolated from DPCs,
form DPCs. Furthermore, heat inactivation of the such as actin, are known to be associated with the
glycosylases prior to incubation with the Oxa substrate nuclear matrix and to be involved in these processes
abolished DPC formation, indicating that the active [19,136–139]. Other proteins, such as the intermediate
form of the protein was needed; however, the same filament protein vimentin, have recently been shown
was not true for histone proteins as heat inactivation to be crosslinked to DNA and to be associated with the
had no effect on DPC formation. These in vitro studies nuclear matrix [7,140]. Because vimentin can bind to
used large excesses of purified proteins and therefore and become crosslinked to DNA, particularly to
may not be biologically relevant, although DPC sequences that resemble sequences at MARs, and
species were also detected (as retarded migration in because it can also bind to histones, it has been
gel shift studies) when the Oxa substrate was proposed that this protein is involved in chromatin
incubated with HeLa cell extract. remodeling [140].
These findings suggest that some types of DNA Cisplatin has been shown to crosslink nuclear
damage are reactive suicide substrates for DNA repair matrix proteins to DNA [18,141]. Nuclear matrix
proteins leading to the further generation of damage fractions and cisplatin-crosslinked fractions were
(i.e., DPCs), and may thereby prevent their own repair. isolated from human breast cancer cells and protein
However, it is not clear if HMG proteins bind damaged profiles were compared by 2-D SDS-PAGE [18,141].
DNA to recruit repair factors, as in the case of HMG Most of the cisplatin-crosslinked proteins were
binding of deoxythioguanosine DNA [131], or bind nuclear matrix proteins. Cisplatin crosslinked several
damaged DNA non-specifically because they recog- transcription factors to the DNA, leading to the
nize any bend in DNA which results in shielding the suggestion that this is a mechanism of transcription
lesion from DNA repair, as is the case for binding of inhibition by crosslinking agents [18]. Additionally,
HMG proteins to cisplatin-modified DNA [132,133]. profiles of crosslinked nuclear matrix proteins
Thus, it is also important to determine which changed in breast cancer cells at different stages of
proteins are responsible for recognizing various types the disease [141].
of DPCs and activating their repair. Clearly, the The effect of IR on the integrity of DNA loop
crosslinking of DNA repair proteins to DNA would be supercoiling was investigated in mouse lymphoma
expected to interfere with the repair process. It may be cells using the propidium iodide fluorescence halo
that repair proteins can become covalently trapped as assay, which allows the visualization of the unwinding
the repairosome moves along the DNA looking for its of anchored DNA loops [135]. The supercoiling
specific lesion substrate. ability of DNA loops was examined in both radio-
S. Barker et al. / Mutation Research 589 (2005) 111–135 125

resistant and radiosensitive cells, with and without the and in mice this frequency increases with age
presence of IR-induced damage, but DPCs were not [148,149]. In mammalian cells, the processes of
specifically analyzed. The supercoiling of DNA loops aging and other cellular stresses (illness, exposure to
containing IR-induced damage was inhibited to a drugs, IR, pollutants, etc.) may result in the
greater degree in radiosensitive cells, suggestive of accumulation of different types of DNA lesions,
alterations in DNA anchoring. This study also used 2- including DPCs, due to oxidative mechanisms
D PAGE to examine the proteins in nucleoids (DNA [148,149]. Nonetheless, the majority of DPCs induced
with associated, extraction-resistant, nuclear matrix by exogenous agents are clearly removed from the
proteins) from both types of cells. Several proteins genome with time (although it should be noted that
associated with nucleoids derived from radioresistant studies of the removal of DPCs from biological
cells were absent from nucleoids from radiosensitive systems are complicated by the known chemical
cells, but none of these proteins correlated directly instability of many types of DPC, as discussed in
with radioresistance [135]. This work provides Section 5). DPCs were detected in rat kidney cells up
evidence of a relationship between IR-induced to 48 h following treatment with nickel compounds
damage and the DNA supercoiling ability of DNA [28]. Levels of ferric nitriloacetate-induced thymine–
loop domains [135]. Balasubramaniam and Oleinick tyrosine DPCs in renal cells of Wistar rats peaked at
[142] demonstrated that IR can crosslink MAR- 24 h (corresponding with the onset of mitosis), but
containing DNA to the nuclear matrix. Clearly, the DPC levels had returned to control level by the 19th
covalent attachment of DNA to the nuclear matrix day of ongoing treatment, suggesting active repair of
should result in serious disregulation of DNA these lesions [56]. Quieveryn and Zhitkovich [24]
metabolic processes. Several studies have indicated reported a half-life for formaldehyde-induced DPCs of
that nuclear matrix proteins are indeed involved in 11.6–13.0 h in three human cell lines (skin, lung, and
DPCs ([19,111,142,143] and others). Stripping his- kidney cell lines) and suggested that the differences in
tones from the DNA with high salt extractions does not DPC half-lives among these cell lines might be due to
completely eliminate the formation of DPCs, indicat- an active repair process. The half-life of formalde-
ing that other proteins, such as nuclear matrix proteins hyde-induced DPCs in peripheral human lymphocytes
that remain bound to DNA despite high salt extraction, was found to be longer (18 h), likely due to
are susceptible to crosslinking by IR [111,112]. inefficient active repair in lymphocytes [150].
Thus, DPC-mediated alterations in the control of Chromium-induced crosslinks were also reported to
DNA supercoiling by altering the anchoring and/or be relatively long lived in human lymphocytes
unwinding of DNA loops might influence DNA repair (reviewed in [74]).
and other processes by altering DNA conformation, At this point, it should be stressed that many of the
remodeling abilities, and/or accessibility. Clearly, the DPC-inducing agents discussed in this review, such as
effect of DPCs on the dynamic control of DNA IR, methylglyoxal, and cisplatin, generate DNA intra-
metabolic processes warrants further investigation. and interstrand crosslinks as well as DPCs. Both DNA
interstrand crosslinks and DPCs are expected to
present special steric challenges to the DNA repair
9. Enzymatic repair of DPCs machinery because of their large size and/or local
covalent involvement of both strands of DNA. Indeed,
Studies on some types of cellular DPCs indicate both of these types of lesions, or at least some sub-
that these lesions can be longer-lived than other types classes thereof, may be repaired by the same pathway
of damage and persist through several DNA-replica- or using some common elements. For example, as will
tion cycles [144,145] and are only partially repaired become apparent, the nucleotide excision repair
[146], which may result in permanent DNA alterations (NER) enzymes ERCC1 and XPF appear to be
and have serious consequences for replication, involved in the repair of some types of DNA-
transcription, and repair processes [147]. A significant interstrand crosslinks as well as DPCs. There are a
background level of accumulated DPCs has been number of outstanding issues in this regard that we
reported in some types of mammalian cells [44,59], will consider in turn.
126 S. Barker et al. / Mutation Research 589 (2005) 111–135

9.1. How are DPCs sensed at the cellular level?

The association of proteins with DNA is a common


occurrence in cellular processes. The mechanisms by
which a cell will distinguish between a protein
associated with DNA appropriately and one that is
bound by a covalent linkage are unknown. Is the DPC
recognized due to its bulk and/or distortion of the
helix? Is the DPC recognized because it blocks the
progression of complexes involved in processes such
as chromatin remodeling, DNA replication, transcrip-
tion, or the repair of other types of lesions?

9.2. How are covalent DPCs repaired?

Depending on the chemistry of the crosslink and


the size and orientation of the protein involved in the
crosslink (i.e., on steric issues), these lesions may be
substrates for different repair pathways. Direct
reversal by chelation (Fig. 5A) is possible in the case
where the protein is bound through complexation with
a metal. Direct reversal by hydrolysis (Fig. 5B) has
been demonstrated for aldehyde-induced DPCs.
At least some DPCs could represent the typical Fig. 5. Potential DPC repair routes: DPCs may be repaired by (A)
direct reversal by chelation, (B) direct reversal by hydrolysis, (C)
bulky/helix distorting adducts that are expected to be
NER, (D) partial proteolytic degradation followed by NER or other
substrates for the NER pathway (Fig. 5C). It may be repair, (E) an incisional–homologous recombinational repair.
that the crosslinking of a protein with extensive DNA
interaction might prevent access to repair enzymes,
and these lesions may first need to be de-bulked by the SDS/K+ precipitation method [24]. Formaldehyde-
proteases before they can be processed by the NER induced DPCs were found to be removed from in vitro
machinery or other repair pathways (Fig. 5D) (see samples by hydrolysis. DPCs in human lung, kidney,
Section 9.4). Alternatively, they may require recom- and fibroblast cells were observed to have a reduced
bination-dependent pathways (Fig. 5E). Several lines half-life compared to formaldehyde-induced DPCs
of evidence suggest that DNA crosslinks are repaired studied in vitro, suggesting that an active repair
through an incisional–recombinational repair mechan- process is involved in DPC loss in cells. Human
ism that involves components of NER and homo- lymphocytes, which are known to have less efficient
logous recombinational repair (HRR) [151,152], NER due to their terminally differentiated status, were
which, in bacteria, is suggested to be the mechanism shown to have reduced DPC removal compared to
involved in restarting stalled replication forks [153]. other human cells studied. However, the human NER-
Previous studies have suggested that there is indeed deficient cell lines, XPA and XPF, were found to have
active repair of DPCs in mammalian cells and that this DPC half-lives similar to that of normal human cells,
may involve more than one repair pathway implying that NER may not be involved in DPC
[33,147,154,155], with NER likely to be involved. removal. Interestingly, XPA cells and, more markedly,
However, the involvement of NER or HRR in DPC XPF cells are sensitive to formaldehyde-induced cell
repair remains unclear. killing. Although formaldehyde induces other types of
The involvement of the NER pathway in removal of DNA damage that are substrates for NER, the
formaldehyde-induced DPCs was examined in several differential sensitivity of the XPA and XPF cells
types of human cells and in vitro with histone H1 using argues for the involvement of the XPF protein in the
S. Barker et al. / Mutation Research 589 (2005) 111–135 127

repair of DNA damage through another pathway. A induced by a transplatin analog, trans-[PtCl2(E-
differential sensitivity to crosslinking agents is also iminoether)2] (trans-EE). Synthetic DPCs were
seen for XPF cells as compared to XPA cells [156], generated by reacting trans-EE with an oligonucleo-
and other studies have suggested the involvement of tide to induce the monoadduct, which was then
the XPF protein but not the XPA protein in a combined with histone H1 to generate the DPC.
recombination-dependent crosslink repair pathway Double-stranded crosslinked substrate was used in in
[157]. vitro reactions to assess the efficiency of incision of
It should be noted, however, that the chemical this lesion by human or rodent cell-free extracts.
instability of many DPCs (see Section 5) means that Incubation of control NER substrates containing a
direct measurements of crosslink repair in different trans-EE-induced monoadduct or a cisplatin-induced
cell types may not be informative for identifying intrastrand crosslink each generated 24–30mer oligo-
proteins involved in the repair of DPCs, and that nucleotide NER excision products, whereas the trans-
studies of the sensitivity of mutant cells to killing by EE-induced DPC substrate showed no excision,
DPC-inducing agents may be more relevant in this indicating that NER is unable to recognize and/or
regard. incise this type of lesion in vitro. These repair assays
Assessments of DPC induction and removal are were performed in vitro, and the protein crosslinking
affected by the limitations of the DPC isolation and was done using single-stranded DNA. It will be of
quantitation method being used. When chromium- interest to see if protein-crosslinking is the predomi-
induced DPCs generated in V79 hamster cells were nant reaction induced by this transplatin analog in vivo
analyzed, no reduced tail moment (i.e., DPCs) was when the complementary DNA strand is present and if
detected [5] by the alkaline comet assay, but a dose- those trans-EE induced lesions are also refractory to
dependent reduction of the tail moment was detected NER.
using a neutral comet assay. The removal of It has recently been shown that the NER system is
formaldehyde-induced DPCs from normal, NER- effective in removing chromium-induced DNA
deficient (XPA), or interstrand crosslink repair- damage. This was quantitated by measuring the initial
deficient (FA-A) human cells was analyzed using and remaining amounts of chromium bound to DNA in
the alkaline comet assay [158]. The XPA and FA-A human XPA, XPC, and XPF cells [160]. Since
cells showed a similar tail moment to the normal cells chromium induces a number of different types of
after formaldehyde treatment, indicating no differ- DNA damage, this study tried to dissect the influence
ences in DPC induction. The tail moments were also of NER specifically on the repair of chromium-
similar for all three formaldehyde-treated cell lines induced crosslinks. Cysteine was crosslinked to a
after various repair times were allowed, indicating no plasmid by chromium treatment and these plasmids
differences in DPC removal between the normal and were transfected into XPA cells and XPA-comple-
repair deficient cell lines. However, there was a dose- mented XPA cells (XPA+ cells). The XPA cells
dependent relationship between formaldehyde con- exhibited significantly greater mutagenic and geno-
centration and the induction of micronuclei in these toxic effects after replication of the crosslink-contain-
human cell lines. The induction of micronuclei might ing plasmid, suggesting the importance of the NER
be due to reduced repair of DPCs and was significantly pathway in dealing with chromium-induced DPCs.
greater in the repair-deficient cell lines, particularly However, this analysis involved only a single amino
the XPA cell line, which argues for a role for NER in acid crosslinked to DNA, and the effect of an entire
the proper repair of DPCs. protein or even a peptide fragment crosslinked to DNA
Alternatively, it may be that chemically distinct may be different.
crosslinks are repaired by different mechanisms and Using a synthetic substrate with an enzyme (T4-
that NER may be involved in the repair of some types pyrimidine dimer–DNA glycosylase) covalently
of DPCs and not others. Transplatin-induced DPCs crosslinked to it, it was shown that the E. coli
have previously been reported to be more persistent in UvrABC complex was capable of incising DNA at the
human XPA cells [33]. A more recent study [159] site of a DPC [161]. Two incisions were made on the
examined the effectiveness of NER in removing DPCs same DNA strand; one incision was made at the eighth
128 S. Barker et al. / Mutation Research 589 (2005) 111–135

phosphodiester bond on the 50 side of the DPC and the hypoxic conditions appear to be distinct from those
second incision was made at the fifth and sixth induced under aerated conditions, which in turn
phosphodiester bonds on the 30 side of the DPC. This influences their repair.
in vitro incision process was more efficient than Several other NER-deficient hamster cell lines,
incisions made on a reduced apurinic/apyrimidinic- notably those with defects in the XPB and XPD genes,
site substrate, but was only half as efficient as that for a did not exhibit this radiosensitive phenotype under
trans-benzo[a]pyrene diol epoxide adduct. The extent hypoxic conditions [100], suggesting that these DPCs
to which this type of repair might be carried out in vivo are not repaired by NER per se. Rather, the
is unclear. phenomenon of hypoxia-specific radiosensitization
Although topoisomerase cleavage complexes are a appears to be restricted to genetic defects that
distinct type of protein–DNA covalent complex, the influence both NER and HRR [100,156], suggesting
mechanisms of their repair may provide some insight that the latter pathway is responsible for the repair of
into the repair of other DPCs. Tyrosyl–DNA DPCs induced by IR in hypoxia. It cannot, however, be
phosphodiesterase I (Tdp1) is an enzyme capable of ruled out that IR-induced DNA interstrand crosslinks
removing a topoisomerase I-covalent complex from a underlie some aspects of these findings since these
DNA end. The activity is specific for hydrolyzing 30 - lesions cannot be realistically measured in irradiated
phosphodiester linkages but could remove a protein mammalian cells at the present time.
other than topoisomerase I from the DNA end
(reviewed in [3]). Tdp1 may also act in strand break 9.4. Might protease activity be involved in DPC
repair pathways as DNA–protein complexes may be repair?
processed into SSBs or DSBs during replication,
transcription, or repair [3]. Cleavage complexes are Proteolytic degradation of the proteins involved in
also substrates for DSB repair pathways, and these DPCs has been suggested to occur in cells. Quievryn
pathways have been shown in yeast to be separate and Zhitkovich [24] demonstrated that formaldehyde-
from the Tdp1 pathway through the use of NER and induced DPCs were removed in part by proteolytic
HRR mutants (reviewed in [3]). degradation because the loss of DPCs was partially
inhibited when cells were incubated with lactacystin, a
9.3. How are IR-induced DPCs repaired? specific inhibitor of proteosomes.
However, cell-cycle regulatory proteins, transcrip-
The involvement of NER in the repair of IR- tion factors and signaling molecules are also
induced DPCs was investigated by assessing the rate substrates for proteolytic degradation; therefore,
of removal of the DPCs induced following irradiation inhibiting proteolysis may affect the induction/
of NER-deficient hamster cells under hypoxic condi- removal/repair of DPCs by mechanisms other than
tions [35]. As measured by alkaline elution, wild-type inhibiting direct proteolytic degradation of the cross-
AA8 cells removed 80% of their DPCs in 24 h, linked protein.
whereas NER-deficient UV41 (XPF) cells removed An earlier study demonstrated that covalent
only 20% of their DPCs in the same period. As was complexes of topoisomerase I and DNA induced by
noted in Section 6.7, UV41 cells are significantly more camptothecin were ubiquitinated and then underwent
sensitive than wild-type AA8 cells to killing by IR proteolytic degradation [163], and this mechanism
under hypoxic conditions, suggesting that a deficiency may be active on other types of DPCs as well.
in the repair of DPCs (which are formed preferentially Proteolytic degradation may not remove the entire
in hypoxia) increases the cells’ radiosensitivity. protein but could leave a small peptide or amino acid
Almost identical cell survival data were reported for adduct, which will then be a substrate for another
the NER-deficient UV20 (ERCC1) hamster cell line, repair pathway, such as NER (Fig. 5C).
although the ability of these cells to repair DPCs was Studies of the Tdp1 enzyme have demonstrated that
not measured [99]. Surprisingly, the repair of DPCs it is more effective on substrates containing a
induced in UV41 cells by IR under aerated conditions denatured or proteolytically digested protein than it
appears to be normal [162]. Thus, DPCs induced under is on substrates containing a native protein [3].
S. Barker et al. / Mutation Research 589 (2005) 111–135 129

Although this enzyme has not been shown to be active Acknowledgements


for other protein–DNA covalent complexes, it is
possible that this or a similarly active enzyme might be This article was made possible through support
involved in a DPC repair step secondary to proteolytic from the Alberta Cancer Board Pilot Project #R-465
digestion. and Research Initiative Program Grant #RI-202 (to
D.M.), and by grants from the Canadian Institutes
of Health Research #MT-14056 (to D.M.) and the
10. Conclusions National Cancer Institute of Canada #013104
with funds from the Canadian Cancer Society (to
Because DPCs have received less attention than M.W.).
other types of DNA damage, their biological
consequences and mechanisms of repair are not well
understood. In part, this is because DPC-inducing References
agents inevitably induce other types of DNA and
protein damage. Possible biochemical consequences [1] G. Neubauer, A. King, J. Rappsilber, C. Calvio, M. Watson, P.
Ajuh, J. Sleeman, A. Lamond, M. Mann, Mass spectrometry
of the covalent crosslinking of proteins to DNA are and EST-database searching allows characterization of the
blockage of replication, transcription and recombina- multi-protein spliceosome complex, Nat. Genet. 20 (1998)
tion. Evidence is mounting that DPCs contribute to the 46–50.
cytotoxic, mutagenic, and carcinogenic effects of a [2] N. Li, A. Mak, D.P. Richards, C. Naber, B.O. Keller, L. Li,
A.R. Shaw, Monocyte lipid rafts contain proteins implicated
number of agents. Further information regarding the
in vesicular trafficking and phagosome formation, Proteomics
mechanisms of the formation and removal of DPCs 3 (2003) 536–548.
would help to delineate the biological relevance of this [3] J.C. Connelly, D.R. Leach, Repair of DNA covalently linked
type of lesion, and may provide insights into cellular to protein, Mol. Cells 13 (2004) 307–316.
processes such as the interaction of the nuclear matrix [4] J.L. Roti Roti, H.H. Kampinga, R.S. Malyapa, W.D. Wright,
with DNA metabolism. Because DPCs are induced by R.P. vanderWaal, M. Xu, Nuclear matrix as a target for
hyperthermic killing of cancer cells, Cell Stress Chaperones
some bifunctional chemotherapy drugs [164,165], the 3 (1998) 245–255.
role of DPCs in the cytotoxicity of these agents may be [5] O. Merk, K. Reiser, G. Speit, Analysis of chromate-induced
relevant to further understanding of clinical responses DNA–protein crosslinks with the comet assay, Mutat. Res.
and drug-resistance mechanisms, which in turn may 471 (2000) 71–80.
lead to novel anti-tumor drug development. For [6] O. Merk, G. Speit, Detection of crosslinks with the comet
assay in relationship to genotoxicity and cytotoxicity,
example, a new analog of transplatin (trans-EE) has Environ. Mol. Mutagen. 33 (1999) 167–172.
been shown to be more cytotoxic than cisplatin and to [7] G.V. Tolstonog, E. Mothes, R.L. Shoeman, P. Traub, Isolation
demonstrate significant anti-tumor activity in both of SDS-stable complexes of the intermediate filament protein
cisplatin-sensitive and -resistant cells [166–168]. vimentin with repetitive, mobile, nuclear matrix attachment
region, and mitochondrial DNA sequence elements from
Compared to cisplatin and cis-EE, the trans-EE
cultured mouse and human fibroblasts, DNA Cell Biol. 20
analog readily crosslinks proteins to DNA, leading to (2001) 531–554.
more efficient inhibition of DNA polymerases and [8] A.E. Cress, K.M. Kurath, M.J. Hendrix, G.T. Bowden,
resistance to NER [159]. Determining the proteins Nuclear protein organization and the repair of radiation
involved in DPCs will also allow us to gain further damage, Carcinogenesis 10 (1989) 939–943.
insight into the consequences of these lesions and their [9] T. Moss, S.I. Dimitrov, D. Houde, UV-laser crosslinking of
proteins to DNA, Methods 11 (1997) 225–234.
repair. The involvement of particular proteins in DPCs [10] S.M. Chiu, N.M. Sokany, L.R. Friedman, N.L. Oleinick,
induced by various environmental and occupational Differential processing of ultraviolet or ionizing radiation-
agents may prove useful in biomonitoring for induced DNA–protein cross-links in Chinese hamster cells,
mutagenesis/carcinogenesis. This review highlights Int. J. Radiat. Biol. Relat. Stud. Phys. Chem. Med. 46 (1984)
the advances in DPC analysis and at the same time 681–690.
[11] L.Y. Xue, L.R. Friedman, N.L. Oleinick, Repair of chromatin
underscores the need for the identification of the damage in glutathione-depleted V-79 cells: comparison of
proteins involved in these lesions and for clarification oxic and hypoxic conditions, Radiat. Res. 116 (1988) 89–
of the mechanisms of their repair. 99.
130 S. Barker et al. / Mutation Research 589 (2005) 111–135

[12] A.E. Cress, K.M. Kurath, B. Stea, G.T. Bowden, The cross- [27] S.K. Chakrabarti, C. Bai, K.S. Subramanian, DNA–protein
linking of nuclear protein to DNA using ionizing radiation, J. crosslinks induced by nickel compounds in isolated rat renal
Cancer Res. Clin. Oncol. 116 (1990) 324–330. cortical cells and its antagonism by specific amino acids and
[13] A. Zhitkovich, M. Costa, A simple, sensitive assay to detect magnesium ion, Toxicol. Appl. Pharmacol. 154 (1999) 245–
DNA–protein crosslinks in intact cells and in vivo, Carcino- 255.
genesis 13 (1992) 1485–1489. [28] S.K. Chakrabarti, C. Bai, K.S. Subramanian, DNA–protein
[14] M. Costa, A. Zhitkovich, M. Gargas, D. Paustenbach, B. crosslinks induced by nickel compounds in isolated rat
Finley, J. Kuykendall, R. Billings, T.J. Carlson, K. Wetter- lymphocytes: role of reactive oxygen species and specific
hahn, J. Xu, S. Patierno, M. Bogdanffy, Interlaboratory amino acids, Toxicol. Appl. Pharmacol. 170 (2001) 153–165.
validation of a new assay for DNA–protein crosslinks, Mutat. [29] A. Zhitkovich, V. Voitkun, M. Costa, Formation of the amino
Res. 369 (1996) 13–21. acid–DNA complexes by hexavalent and trivalent chromium
[15] K.W. Kohn, DNA as a target in cancer chemotherapy: mea- in vitro: importance of trivalent chromium and the phosphate
surement of macromolecular DNA damage produced in group, Biochemistry 35 (1996) 7275–7282.
mammalian cells by anticancer agents and carcinogens, in: [30] A. Mozumder, Early production of radicals from charged
Methods in Cancer Research, Academic Press, New York, particle tracks in water, Radiat. Res. Suppl. 8 (1985) S33–S39.
1979, pp. 291–345. [31] E.C. Woudstra, Chromatin Structure, DNA Damage, DNA
[16] K.W. Kohn, R.A. Ewig, L.C. Erickson, L.A. Zwelling, Mea- Repair and Cellular Radiosensitivity, Radiobiology, Ph.D.
surement of strand breaks and crosslinks by alkaline elution, Thesis, University of Groningen, The Netherlands, 1999,
in: E.C. Friedberg, P.C. Hanawalt (Eds.), DNA Repair: A p. 270.
Laboratory Manual of Research Procedures, Marcel Dekker, [32] A.J. Fornace Jr., J.B. Little, DNA crosslinking induced by X-
New York, 1981, pp. 379–401. rays and chemical agents, Biochim. Biophys. Acta 477 (1977)
[17] A. Ferraro, P. Grandi, M. Eufemi, F. Altieri, C. Turano, 343–355.
Crosslinking of nuclear proteins to DNA by cis-diamminedi- [33] A.J. Fornace Jr., J.B. Little, DNA–protein cross-linking by
chloroplatinum in intact cells. Involvement of nuclear matrix chemical carcinogens in mammalian cells, Cancer Res. 39
proteins, FEBS Lett. 307 (1992) 383–385. (1979) 704–710.
[18] S.K. Samuel, V.A. Spencer, L. Bajno, J.M. Sun, L.T. Holth, S. [34] R.E. Meyn, W.T. Jenkins, D. Murray, Radiation damage to
Oesterreich, J.R. Davie, In situ cross-linking by cisplatin of DNA in various animal tissues: comparison of yields and
nuclear matrix-bound transcription factors to nuclear DNA of repair in vivo and in vitro, in: M.G. Simic, L. Grossman, A.C.
human breast cancer cells, Cancer Res. 58 (1998) 3004–3008. Upton (Eds.), Mechanisms of DNA Damage and Repair:
[19] C.A.I. Miller, M.D. Cohen, M. Costa, Complexing of actin Implications to Carcinogenesis and Risk Assessment, Plenum
and other nuclear proteins to DNA by cis-diamminedichlor- Press, New York, 1986, pp. 151–158.
oplatinum(II) and chromium compounds, Carcinogenesis 12 [35] R.E. Meyn, S.C. vanAnkeren, W.T. Jenkins, The induction of
(1991) 269–276. DNA–protein crosslinks in hypoxic cells and their possible
[20] A.E. Cress, K.M. Kurath, Identification of attachment pro- contribution to cell lethality, Radiat. Res. 109 (1987) 419–
teins for DNA in Chinese hamster ovary cells, J. Biol. Chem. 429.
263 (1988) 19678–19683. [36] I.R. Radford, Effect of radiomodifying agents on the ratios of
[21] S.N. Mattagajasingh, H.P. Misra, Mechanisms of the carci- X-ray-induced lesions in cellular DNA: use in lethal lesion
nogenic chromium(VI)-induced DNA–protein cross-linking determination, Int. J. Radiat. Biol. Relat. Stud. Phys. Chem.
and their characterization in cultured intact human cells, J. Med. 49 (1986) 621–637.
Biol. Chem. 271 (1996) 33550–33560. [37] H. Zhang, K.T. Wheeler, Radiation-induced DNA damage in
[22] S.N. Mattagajasingh, H.P. Misra, Analysis of EDTA-chela- tumors and normal tissues. I. Feasibility of estimating the
table proteins from DNA–protein crosslinks induced by a hypoxic fraction, Radiat. Res. 136 (1993) 77–88.
carcinogenic chromium (VI) in cultured intact human cells, [38] H. Zhang, K.T. Wheeler, Radiation-induced DNA damage in
Mol. Cell. Biochem. 199 (1999) 149–162. tumors and normal tissues. II. Influence of dose, residual
[23] T.H. Ma, M.M. Harris, Review of the genotoxicity of for- DNA damage and physiological factors in oxygenated cells,
maldehyde, Mutat. Res. 196 (1988) 37–59. Radiat. Res. 140 (1994) 321–326.
[24] G. Quievryn, A. Zhitkovich, Loss of DNA–protein crosslinks [39] H. Zhang, C.J. Koch, C.A. Wallen, K.T. Wheeler, Radiation-
from formaldehyde-exposed cells occurs through sponta- induced DNA damage in tumors and normal tissues. III.
neous hydrolysis and an active repair process linked to Oxygen dependence of the formation of strand breaks and
proteosome function, Carcinogenesis 21 (2000) 1573– DNA–protein crosslinks, Radiat. Res. 142 (1995) 163–168.
1580. [40] I. Al-Nabulsi, K.T. Wheeler, Temperature dependence of
[25] J.D. McGhee, P.H. von Hippel, Formaldehyde as a probe of radiation-induced DNA–protein crosslinks formed under
DNA structure. II. Reaction with endocyclic imino groups of hypoxic conditions, Radiat. Res. 148 (1997) 568–574.
DNA bases, Biochemistry 14 (1975) 1297–1303. [41] L.K. Mee, S.J. Adelstein, Predominance of core histones in
[26] J.D. McGhee, P.H. von Hippel, Formaldehyde as a probe of formation of DNA–protein crosslinks in gamma-irradiated
DNA structure. I. Reaction with exocyclic amino groups of chromatin, Proc. Natl. Acad. Sci. U.S.A. 78 (1981) 2194–
DNA bases, Biochemistry 14 (1975) 1281–1296. 2198.
S. Barker et al. / Mutation Research 589 (2005) 111–135 131

[42] A.E. Cress, G.T. Bowden, Covalent DNA–protein crosslink- tyrosine in their renal chromatin, Int. J. Cancer 62 (1995)
ing occurs after hyperthermia and radiation, Radiat. Res. 95 309–313.
(1983) 610–619. [57] M.S. Weir Lipton, A.F. Fuciarelli, D.L. Springer, C.G.
[43] S.M. Chiu, L.R. Friedman, N.L. Oleinick, Formation and Edmonds, Characterization of radiation-induced thymine–
repair of DNA–protein crosslinks in newly replicated DNA, tyrosine crosslinks by electrospray ionization mass spectro-
Radiat. Res. 120 (1989) 545–551. metry, Radiat. Res. 145 (1996) 681–686.
[44] R. Olinski, Z. Nackerdien, M. Dizdaroglu, DNA–protein [58] S. Gebicki, J.M. Gebicki, Crosslinking of DNA and proteins
cross-linking between thymine and tyrosine in chromatin induced by protein hydroperoxides, Biochem. J. 338 (1999)
of gamma-irradiated or H2O2-treated cultured human cells, 629–636.
Arch. Biochem. Biophys. 297 (1992) 139–143. [59] V. Voitkun, A. Zhitkovich, Analysis of DNA–protein cross-
[45] I. Al-Nabulsi, K.T. Wheeler, Radiation-induced DNA linking activity of malondialdehyde in vitro, Mutat. Res. 424
damage in tumors and normal tissues: V. Influence of pH (1999) 97–106.
and nutrient depletion on the formation of DNA–protein [60] C.L. Hawkins, D.I. Pattison, M.J. Davies, Reaction of protein
crosslinks in irradiated partially and fully hypoxic tumor chloramines with DNA and nucleosides: evidence for the
cells, Radiat. Res. 151 (1999) 188–194. formation of radicals, protein–DNA cross-links and DNA
[46] J.F. Ward, Biochemistry of DNA lesions, Radiat. Res. Suppl. fragmentation, Biochem. J. 365 (2002) 605–615.
8 (1985) S103–S111. [61] J.R. Kuykendall, M.S. Bogdanffy, Reaction kinetics of DNA–
[47] H. Nikjoo, P. O’Neill, M. Terrissol, D.T. Goodhead, Model- histone crosslinking by vinyl acetate and acetaldehyde, Car-
ling of radiation-induced DNA damage: the early physical cinogenesis 13 (1992) 2095–2100.
and chemical event, Int. J. Radiat. Biol. 66 (1994) 453–457. [62] M. Costa, A. Zhitkovich, M. Harris, D. Paustenbach, M.
[48] D.T. Goodhead, Molecular and cell models of biological Gargas, DNA–protein cross-links produced by various che-
effects of heavy ion radiation, Radiat. Environ. Biophys. micals in cultured human lymphoma cells, J. Toxicol.
34 (1995) 67–72. Environ. Health 50 (1997) 433–449.
[49] E.A. Blakely, P.Y. Chang, K.A. Bjornstad, LET dependence [63] P. Bedinger, M. Hochstrasser, C.V. Jongeneel, B.M. Alberts,
of DNA–protein crosslinks, in: U. Hagen, H. Jung, C. Streffer Properties of the T4 bacteriophage DNA replication appara-
(Eds.), Radiation Research 1895–1995, Congress Proceed- tus: the T4 dda DNA helicase is required to pass a bound RNA
ings, ICRR Society, 1995, pp. 136–139. polymerase molecule, Cell 34 (1983) 115–123.
[50] K. Eguchi, T. Inada, M. Yaguchi, S. Satoh, I. Kaneko, [64] A.L. Pinto, S.J. Lippard, Sequence-dependent termination of
Induction and repair of DNA lesions in cultured human in vitro DNA synthesis by cis- and trans-diamminedichloro-
melanoma cells exposed to a nitrogen-ion beam, Int. J. platinum (II), Proc. Natl. Acad. Sci. U.S.A. 82 (1985) 4616–
Radiat. Biol. Relat. Stud. Phys. Chem. Med. 52 (1987) 4619.
115–123. [65] J.A. Briggs, R.C. Briggs, Characterization of chromium
[51] M. Frankenberg-Schwager, Radiation-induced DNA lesions effects on a rat liver epithelial cell line and their relevance
in eukaryotic cells, their repair and biological relevance, in: to in vitro transformation, Cancer Res. 48 (1988) 6484–6490.
C.E. Swenberg, G. Horneck, E.G. Stassinopoulos (Eds.), [66] M.O. Bradley, I.C. Hsu, C.C. Harris, Relationship between
Biological Effects and Physics of Solar and Galactic Cosmic sister chromatid exchange and mutagenicity, toxicity and
Radiation, Plenum Press, New York, 1993, pp. 1–31. DNA damage, Nature 282 (1979) 318–320.
[52] E. Gajewski, A.F. Fuciarelli, M. Dizdaroglu, Structure of [67] M.O. Bradley, K.W. Kohn, X-ray induced DNA double strand
hydroxyl radical-induced DNA–protein crosslinks in calf break production and repair in mammalian cells as measured
thymus nucleohistone in vitro, Int. J. Radiat. Biol. 54 by neutral filter elution, Nucleic Acids Res. 7 (1979) 793–804.
(1988) 445–459. [68] A.J. Fornace Jr., J.B. Little, Malignant transformation by the
[53] M. Dizdaroglu, E. Gajewski, Structure and mechanism of DNA–protein crosslinking agent trans-Pt(II) diammine-
hydroxyl radical-induced formation of a DNA–protein cross- dichloride, Carcinogenesis 1 (1980) 989–994.
link involving thymine and lysine in nucleohistone, Cancer [69] A.J. Fornace Jr., Detection of DNA single-strand breaks
Res. 49 (1989) 3463–3467. produced during the repair of damage by DNA–protein cross-
[54] M. Dizdaroglu, E. Gajewski, P. Reddy, S.A. Margolis, Struc- linking agents, Cancer Res. 42 (1982) 145–149.
ture of a hydroxyl radical induced DNA–protein cross-link [70] O. Merk, G. Speit, Significance of formaldehyde-induced
involving thymine and tyrosine in nucleohistone, Biochem- DNA–protein crosslinks for mutagenesis, Environ. Mol.
istry 28 (1989) 3625–3628. Mutagen. 32 (1998) 260–268.
[55] S.A. Altman, T.H. Zastawny, L. Randers-Eichhorn, M.A. [71] H. Waksvik, M. Boysen, Cytogenetic analyses of lympho-
Cacciuttolo, S.A. Akman, M. Dizdaroglu, G. Rao, Formation cytes from workers in a nickel refinery, Mutat. Res. 103
of DNA–protein cross-links in cultured mammalian cells (1982) 185–190.
upon treatment with iron ions, Free Radic. Biol. Med. 19 [72] H. Waksvik, M. Boysen, A.C. Hogetveit, Increased incidence
(1995) 897–902. of chromosomal aberrations in peripheral lymphocytes of
[56] S. Toyokuni, T. Mori, H. Hiai, M. Dizdaroglu, Treatment of retired nickel workers, Carcinogenesis 5 (1984) 1525–1527.
Wistar rats with a renal carcinogen, ferric nitrilotriacetate, [73] J.K. Dunnick, M.R. Elwell, A.E. Radovsky, J.M. Benson, F.F.
causes DNA–protein cross-linking between thymine and Hahn, K.J. Nikula, E.B. Barr, C.H. Hobbs, Comparative
132 S. Barker et al. / Mutation Research 589 (2005) 111–135

carcinogenic effects of nickel subsulfide, nickel oxide, or with DNA–protein cross-link measurements in the rat nasal
nickel sulfate hexahydrate chronic exposures in the lung, passages, Toxicol. Appl. Pharmacol. 143 (1997) 47–55.
Cancer Res. 55 (1995) 5251–5256. [89] M. Casanova, K.T. Morgan, W.H. Steinhagen, J.I. Everitt,
[74] A. Zhitkovich, V. Voitkun, T. Kluz, M. Costa, Utilization of J.A. Popp, H.D. Heck, Covalent binding of inhaled formal-
DNA–protein cross-links as a biomarker of chromium expo- dehyde to DNA in the respiratory tract of rhesus monkeys:
sure, Environ. Health Perspect. 106 (Suppl. 4) (1998) 969– pharmacokinetics, rat-to-monkey interspecies scaling, and
974. extrapolation to man, Fundam. Appl. Toxicol. 17 (1991)
[75] V. Voitkun, A. Zhitkovich, M. Costa, Cr(III)-mediated cross- 409–428.
links of glutathione or amino acids to the DNA phosphate [90] M. Casanova, K.T. Morgan, E.A. Gross, O.R. Moss, H.A.
backbone are mutagenic in human cells, Nucleic Acids Res. Heck, DNA–protein cross-links and cell replication at spe-
26 (1998) 2024–2030. cific sites in the nose of F344 rats exposed subchronically to
[76] M.N. Bates, A.H. Smith, C. Hopenhayn-Rich, Arsenic inges- formaldehyde, Fundam. Appl. Toxicol. 23 (1994) 525–536.
tion and internal cancers: a review, Am. J. Epidemiol. 135 [91] N. Murata-Kamiya, H. Kamiya, H. Kaji, H. Kasai, Methyl-
(1992) 462–476. glyoxal induces G:C to C:G and G:C to T:A transversions in
[77] C.J. Chen, C.W. Chen, M.M. Wu, T.L. Kuo, Cancer potential the supF gene on a shuttle vector plasmid replicated in
in liver, lung, bladder and kidney due to ingested inorganic mammalian cells, Mutat. Res. 468 (2000) 173–182.
arsenic in drinking water, Br. J. Cancer 66 (1992) 888–892. [92] B.M. Barnett, E.R. Munoz, Genetic damage induced by
[78] M.E. Gonsebatt, L. Vega, A.M. Salazar, R. Montero, P. methylglyoxal and methylglyoxal plus X-rays in Drosophila
Guzman, J. Blas, L.M. Del Razo, G. Garcia-Vargas, A. melanogaster germinal cells, Mutat. Res. 421 (1998) 37–43.
Albores, M.E. Cebrian, M. Kelsh, P. Ostrosky-Wegman, [93] N. Murata-Kamiya, H. Kamiya, Methylglyoxal, an endogen-
Cytogenetic effects in human exposure to arsenic, Mutat. ous aldehyde, crosslinks DNA polymerase and the substrate
Res. 386 (1997) 219–228. DNA, Nucleic Acids Res. 29 (2001) 3433–3438.
[79] D.T. Bau, T.S. Wang, C.H. Chung, A.S. Wang, K.Y. Jan, [94] M.J. Roberts, G.T. Wondrak, D.C. Laurean, M.K. Jacobson,
Oxidative DNA adducts and DNA–protein cross-links are the E.L. Jacobson, DNA damage by carbonyl stress in human
major DNA lesions induced by arsenite, Environ. Health skin cells, Mutat. Res. 522 (2003) 45–56.
Perspect. 110 (Suppl. 5) (2002) 753–756. [95] R.J. Huxtable, Human health implications of pyrrolizidine
[80] S. Lynn, J.R. Gurr, H.T. Lai, K.Y. Jan, NADH oxidase alkaloids and herbs containing them, in: P.R. Cheeke (Ed.),
activation is involved in arsenite-induced oxidative DNA Toxicants of Plant Origin, vol. I. Alkaloids, CRC Press Inc,
damage in human vascular smooth muscle cells, Circ. Res. Boca Raton, FL, 1989, pp. 41–86.
86 (2000) 514–519. [96] J.R. Hincks, H.Y. Kim, H.J. Segall, R.J. Molyneux, F.R.
[81] M. Whiteman, B. Halliwell, Loss of 3-nitrotyrosine on expo- Stermitz, R.A. Coulombe Jr., DNA cross-linking in mam-
sure to hypochlorous acid: implications for the use of 3- malian cells by pyrrolizidine alkaloids: structure–activity
nitrotyrosine as a bio-marker in vivo, Biochem. Biophys. Res. relationships, Toxicol. Appl. Pharmacol. 111 (1991) 90–98.
Commun. 258 (1999) 168–172. [97] H.Y. Kim, F.R. Stermitz, R.A. Coulombe Jr., Pyrrolizidine
[82] M. Whiteman, J.P. Spencer, A. Jenner, B. Halliwell, Hypo- alkaloid-induced DNA–protein cross-links, Carcinogenesis
chlorous acid-induced DNA base modification: potentiation by 16 (1995) 2691–2697.
nitrite: biomarkers of DNA damage by reactive oxygen species, [98] H.Y. Kim, F.R. Stermitz, J.K. Li, R.A. Coulombe Jr., Com-
Biochem. Biophys. Res. Commun. 257 (1999) 572–576. parative DNA cross-linking by activated pyrrolizidine alka-
[83] N. Mei, J. Lee, X. Sun, J.Z. Xing, J. Hanson, X.C. Le, M. loids, Food Chem. Toxicol. 37 (1999) 619–625.
Weinfeld, Genetic predisposition to the cytotoxicity of [99] D. Murray, A. Macann, J. Hanson, E. Rosenberg, ERCC1/
arsenic: the role of DNA damage and ATM, FASEB J. 17 ERCC4 50 -endonuclease activity as a determinant of hypoxic
(2003) 2310–2312. cell radiosensitivity, Int. J. Radiat. Biol. 69 (1996) 319–327.
[84] G. Speit, O. Merk, Evaluation of mutagenic effects of for- [100] D. Murray, E. Rosenberg, The importance of the ERCC1/
maldehyde in vitro: detection of crosslinks and mutations in ERCC4[XPF] complex for hypoxic-cell radioresistance does
mouse lymphoma cells, Mutagenesis 17 (2002) 183–187. not appear to derive from its participation in the nucleotide
[85] W.D. Kerns, K.L. Pavkov, D.J. Donofrio, E.J. Gralla, J.A. excision repair pathway, Mutat. Res. 364 (1996) 217–226.
Swenberg, Carcinogenicity of formaldehyde in rats and mice [101] F.Y. Wu, Y.J. Lee, D.R. Chen, H.W. Kuo, Association of
after long-term inhalation exposure, Cancer Res. 43 (1983) DNA–protein crosslinks and breast cancer, Mutat. Res. 501
4382–4392. (2002) 69–78.
[86] L. Recio, S. Sisk, L. Pluta, E. Bermudez, E.A. Gross, Z. Chen, [102] A. Matsumoto, P.C. Hanawalt, Histone H3 and heat shock
K. Morgan, C. Walker, p53 mutations in formaldehyde- protein GRP78 are selectively cross-linked to DNA by photo-
induced nasal squamous cell carcinomas in rats, Cancer activated Gilvocarcin V in human fibroblasts, Cancer Res. 60
Res. 52 (1992) 6113–6116. (2000) 3921–3926.
[87] USEPA Formaldehyde Risk Assessment Update, Office of [103] P. Ramirez, L.M. Del Razo, M.C. Gutierrex-Ruiz, M.E.
Toxic Substances, Washington, DC, 1991. Gonsebatt, Arsenite induces DNA–protein crosslinks and
[88] E.A. Hubal, P.M. Schlosser, R.B. Conolly, J.S. Kimbell, cytokeratin expression in the WRL-68 human hepatic cell
Comparison of inhaled formaldehyde dosimetry predictions line, Carcinogenesis 21 (2000) 701–706.
S. Barker et al. / Mutation Research 589 (2005) 111–135 133

[104] P.M. O’Connor, B.W. Fox, Isolation and characterization of [121] L. Beckerbauer, J.J. Tepe, J. Cullison, R. Reeves, R.M.
proteins cross-linked to DNA by the antitumor agent methy- Williams, FR900482 class of anti-tumor drugs cross-links
lene dimethanesulfonate and its hydrolytic product formal- oncoprotein HMG I/Y to DNA in vivo, Chem. Biol. 7 (2000)
dehyde, J. Biol. Chem. 264 (1989) 6391–6397. 805–812.
[105] R.P. VanderWaal, D.R. Spitz, C.L. Griffith, R. Higashikubo, [122] J.F. Wang, M. Bashir, B.N. Engelsberg, C. Witmer, H.
J.L. Roti Roti, Evidence that protein disulfide isomerase Rozmiarek, P.C. Billings, High mobility group proteins 1
(PDI) is involved in DNA–nuclear matrix anchoring, J. Cell. and 2 recognize chromium-damaged DNA, Carcinogenesis
Biochem. 85 (2002) 689–702. 18 (1997) 371–375.
[106] R.A.J. Coulombe, G.L. Drew, F.R. Stermitz, Pyrrolizidine [123] P.C. Billings, R.J. Davis, B.N. Engelsberg, K.A. Skov, E.N.
alkaloids crosslink DNA with actin, Toxicol. Appl. Pharma- Hughes, Characterization of high mobility group protein
col. 15 (1999) 198–202. binding to cisplatin-damaged DNA, Biochem. Biophys.
[107] B. Singer, D. Grunberger, Reactions of directly acting agents Res. Commun. 188 (1992) 1286–1294.
with nucleic acids, in: B. Singer, D. Grunberger (Eds.), [124] P.A. Kulcharyk, J.W. Heinecke, Hypochlorous acid produced
Molecular Biology of Mutagens and Carcinogens, Plenum by the myeloperoxidase system of human phagocytes induces
Publishing Corp., New York, 1983, pp. 53–55. covalent cross-links between DNA and protein, Biochemistry
[108] C.A. Miller III, M. Costa, Analysis of proteins cross-linked to 40 (2001) 3648–3656.
DNA after treatment of cells with formaldehyde, chromate, [125] T.A. Rosenquist, E. Zaika, A.S. Fernandes, D.O. Zharkov, H.
and cis-diamminedichloroplatinum(II), Mol. Toxicol. 2 Miller, A.P. Grollman, The novel DNA glycosylase, NEIL1,
(1989) 11–26. protects mammalian cells from radiation-mediated cell death,
[109] J.R. Kuykendall, M.S. Bogdanffy, Efficiency of DNA–his- DNA Repair (Amst.) 2 (2003) 581–591.
tone crosslinking induced by saturated and unsaturated alde- [126] M. Hashimoto, M.M. Greenberg, Y.W. Kow, J.-T. Hwang,
hydes in vitro, Mutat. Res. 283 (1992) 131–136. R.P. Cunningham, The 2-deoxyribonolactone lesion pro-
[110] J.R. Kuykendall, M.S. Bogdanffy, Formation and stability of duced in DNA by neocarzinostatin and other damaging agents
acetaldehyde-induced crosslinks between poly-lysine and forms cross-links with the base-excision repair enzyme endo-
poly-deoxyguanosine, Mutat. Res. 311 (1994) 49–56. nuclease III, J. Am. Chem. Soc. 123 (2001) 3161–3162.
[111] L.Y. Xue, L.R. Friedman, N.L. Oleinick, S.M. Chiu, Induc- [127] M.S. DeMott, E. Beyret, D. Wong, B.C. Bales, J.T. Hwang,
tion of DNA damage in gamma-irradiated nuclei stripped of M.M. Greenberg, B. Demple, Covalent trapping of human
nuclear protein classes: differential modulation of double- DNA polymerase beta by the oxidative DNA lesion 2-deox-
strand break and DNA–protein crosslink formation, Int. J. yribonolactone, J. Biol. Chem. 277 (2002) 7637–7640.
Radiat. Biol. 66 (1994) 11–21. [128] T. Suzuki, M. Yamada, H. Ide, K. Kanaori, K. Tajima, T.
[112] N.L. Oleinick, U. Balasubramaniam, L. Xue, S. Chiu, Morii, K. Makino, Identification and characterization of a
Nuclear structure and the microdistribution of radiation reaction product of 20 -deoxyoxanosine with glycine, Chem.
damage in DNA, Int. J. Radiat. Biol. 66 (1994) 523–529. Res. Toxicol. 13 (2000) 227–230.
[113] A. Agresti, M.E. Bianchi, HMGB proteins and gene expres- [129] T. Suzuki, M. Yamada, T. Nakamura, H. Ide, K. Kanaori, K.
sion, Curr. Opin. Genet. Dev. 13 (2003) 170–178. Tajima, T. Morii, K. Makino, Products of the reaction
[114] M. Bustin, HMGN dynamics and chromatin function, Bio- between a diazoate derivative of 20 -deoxycytidine and L-
chem. Cell Biol. 81 (2003) 113–122. lysine and its implication for DNA–nucleoprotein cross-
[115] R. Reeves, HMGA proteins: flexibility finds a nuclear niche? linking by NO or HNO(2), Chem. Res. Toxicol. 13 (2000)
Biochem. Cell Biol. 81 (2003) 185–195. 1223–1227.
[116] A.A. Travers, Priming the nucleosome: a role for HMGB [130] T. Suzuki, T. Nakamura, M. Yamada, H. Ide, K. Kanaori, K.
proteins? EMBO Rep. 4 (2003) 131–136. Tajima, T. Morii, K. Makino, Isolation and characterization of
[117] T. Nakano, H. Terato, K. Asagoshi, A. Masaoka, M. Mukuta, diazoate intermediate upon nitrous acid and nitric oxide
Y. Ohyama, T. Suzuki, K. Makino, H. Ide, DNA–protein treatment of 20 -deoxycytidine, Biochemistry 38 (1999)
cross-link formation mediated by oxanine. A novel genotoxic 7151–7158.
mechanism of nitric oxide-induced DNA damage, J. Biol. [131] E.Y. Krynetski, N.F. Krynetskaia, M.E. Bianchi, W.E. Evans,
Chem. 278 (2003) 25264–25272. A nuclear protein complex containing high mobility group
[118] S.R. Rajski, S.B. Rollins, R.M. Williams, FR-66979 cova- proteins B1 and B2, heat shock cognate protein 70, ERp60,
lently cross-links the binding domain of the high-mobility and glyceraldehyde-3-phosphate dehydrogenase is involved
group I/Y proteins to DNA, J. Am. Chem. Soc. 120 (1998) in the cytotoxic response to DNA modified by incorporation
2192–2193. of anticancer nucleoside analogues, Cancer Res. 63 (2003)
[119] S.R. Rajski, R.M. Williams, Observations on the covalent 100–106.
cross-linking of the binding domain (BD) of the high mobility [132] Q. He, C.H. Liang, S.J. Lippard, Steroid hormones induce
group I/Y (HMG I/Y) proteins to DNA by FR66979, Bioorg. HMG1 overexpression and sensitize breast cancer cells to
Med. Chem. 8 (2000) 1331–1342. cisplatin and carboplatin, Proc. Natl. Acad. Sci. U.S.A. 97
[120] M. Bustin, Revised nomenclature for high mobility group (2000) 5768–5772.
(HMG) chromosomal proteins, Trends Biochem. Sci. 26 [133] G. Nagatani, M. Nomoto, H. Takano, T. Ise, K. Kato, T.
(2001) 152–153. Imamura, H. Izumi, K. Makishima, K. Kohno, Transcrip-
134 S. Barker et al. / Mutation Research 589 (2005) 111–135

tional activation of the human HMG1 gene in cisplatin- cross-links in mammalian cells, Br. J. Cancer Suppl. 8 (1987)
resistant human cancer cells, Cancer Res. 61 (2001) 1592– 135–140.
1597. [148] A. Izzotti, C. Cartiglia, M. Taningher, S. De Flora, R.
[134] E.S. Leman, R.H. Betzenberg, Nuclear matrix proteins as Balansky, Age-related increases of 8-hydroxy-20 -deoxygua-
biomarkers in prostate cancer, J. Cell. Biochem. 86 (2002) nosine and DNA–protein crosslinks in mouse organs, Mutat.
213–223. Res. 446 (1999) 215–223.
[135] R.S. Malyapa, W.D. Wright, J.L. Roti Roti, DNA super- [149] R.K. Zahn, G. Zahn-Daimler, S. Ax, M. Hosokawa, T.
coiling changes and nucleoid protein composition in a group Takeda, Assessment of DNA–protein crosslinks in the course
of L5178Y cells of varying radiosensitivity, Radiat. Res. 145 of aging in two mouse strains by use of a modified alkaline
(1996) 239–242. filter elution applied to whole tissue samples, Mech. Ageing
[136] D. Rungger, E. Rungger-Brandle, C. Chaponnier, G. Gab- Dev. 108 (1999) 99–112.
biani, Intranuclear injection of anti-actin antibodies into [150] J.M. Barret, P. Calsou, B. Salles, Deficient nucleotide exci-
Xenopus oocytes blocks chromosome condensation, Nature sion repair activity in protein extracts from normal human
282 (1979) 320–321. lymphocytes, Carcinogenesis 16 (1995) 1611–1616.
[137] J.M. Egly, N.G. Miyamoto, V. Moncollin, P. Chambon, Is [151] L.H. Thompson, Evidence that mammalian cells possess
actin a transcription initiation factor for RNA polymerase B? homologous recombinational repair pathways, Mutat. Res.
EMBO J. 3 (1984) 2363–2371. 363 (1996) 77–88.
[138] J.W. Hamilton, C.A. Louis, K.A. Doherty, S.R. Hunt, M.J. [152] R.J. Legerski, C. Richie, Mechanisms of repair of interstrand
Reed, M.D. Treadwell, Preferential alteration of inducible crosslinks in DNA, Cancer Treat. Res. 112 (2002) 109–
gene expression in vivo by carcinogens that induce bulky 128.
DNA lesions, Mol. Carcinog. 8 (1993) 34–43. [153] M.M. Cox, The nonmutagenic repair of broken replication
[139] W. Hofmann, B. Reichart, A. Ewald, E. Muller, I. Schmitt, forks via recombination, Mutat. Res. 510 (2002) 107–120.
R.H. Stauber, F. Lottspeich, B.M. Jockusch, U. Scheer, J. [154] R.C. Grafstrom, A. Fornace Jr., C.C. Harris, Repair of DNA
Hauber, M.C. Dabauvalle, Cofactor requirements for nuclear damage caused by formaldehyde in human cells, Cancer Res.
export of Rev response element (RRE)- and constitutive 44 (1984) 4323–4327.
transport element (CTE)-containing retroviral RNAs. An [155] R. Gantt, A cell cycle-associated pathway for repair of DNA–
unexpected role for actin, J. Cell Biol. 152 (2001) 895–910. protein crosslinks in mammalian cells, Mutat. Res. 183
[140] G.V. Tolstonog, M. Sabasch, P. Traub, Cytoplasmic inter- (1987) 75–87.
mediate filaments are stably associated with nuclear matrices [156] D. Murray, L. Vallee-Lucic, E. Rosenberg, B. Andersson,
and potentially modulate their DNA-binding function, DNA Sensitivity of nucleotide excision repair-deficient human
Cell Biol. 21 (2002) 213–239. cells to ionizing radiation and cyclophosphamide, Anticancer
[141] V.A. Spencer, S.K. Samuel, J.R. Davie, Altered profiles in Res. 22 (2002) 21–26.
nuclear matrix proteins associated with DNA in situ during [157] L. Li, C.A. Peterson, X. Lu, P. Wei, R.J. Legerski, Interstrand
progression of breast cancer cells, Cancer Res. 61 (2001) cross-links induce DNA synthesis in damaged and unda-
1362–1366. maged plasmids in mammalian cell extracts, Mol. Cell. Biol.
[142] U. Balasubramaniam, N.L. Oleinick, Preferential cross-link- 19 (1999) 5619–5630.
ing of matrix-attachment region (MAR) containing DNA [158] G. Speit, P. Schutz, O. Merk, Induction and repair of for-
fragments to the isolated nuclear matrix by ionizing radiation, maldehyde-induced DNA–protein crosslinks in repair-defi-
Biochemistry 34 (1995) 12790–12802. cient human cell lines, Mutagenesis 15 (2000) 85–90.
[143] N.L. Oleinick, S.M. Chiu, L.R. Friedman, L.Y. Xue, N. [159] O. Novakova, J. Kasparkova, J. Malina, G. Natile, V. Brabec,
Ramakrishnan, DNA–protein cross-links: new insights into DNA–protein cross-linking by trans-[PtCl(2)(E-imi-
their formation and repair in irradiated mammalian cells, noether)(2)]. A concept for activation of the trans geometry
Basic Life Sci. 38 (1986) 181–192. in platinum antitumor complexes, Nucleic Acids Res. 31
[144] M.J. Tsapakos, T.H. Hampton, K.E. Wetterhahn, Chro- (2003) 6450–6460.
mium(VI)-induced DNA lesions and chromium distribution [160] M. Reynolds, E. Peterson, G. Quievryn, A. Zhitkovich,
in rat kidney, liver, and lung, Cancer Res. 43 (1983) 5662– Human nucleotide excision repair efficiently removes chro-
5667. mium–DNA phosphate adducts and protects cells against
[145] D.Y. Cupo, K.E. Wetterhahn, Binding of chromium to chro- chromate toxicity, J. Biol. Chem. 279 (2004) 30419–30424.
matin and DNA from liver and kidney of rats treated with [161] I.G. Minko, Y. Zou, R.S. Lloyd, Incision of DNA–protein
sodium dichromate and chromium(III) chloride in vivo, crosslinks by UvrABC nuclease suggests a potential repair
Cancer Res. 45 (1985) 1146–1151. pathway involving nucleotide excision repair, Proc. Natl.
[146] M. Sugiyama, S.R. Patierno, O. Cantoni, M. Costa, Char- Acad. Sci. U.S.A. 99 (2002) 1905–1909.
acterization of DNA lesions induced by CaCrO4 in synchro- [162] S.-M. Chiu, L.-Y. Xue, L.R. Friedman, N.L. Oleinick, DNA–
nous and asynchronous cultured mammalian cells, Mol. protein crosslinks: formation and repair in irradiated cells, in:
Pharmacol. 29 (1986) 606–613. U. Hagen, H. Jung, C. Streffer (Eds.), Radiation Research
[147] N.L. Oleinick, S.M. Chiu, N. Ramakrishnan, L.Y. Xue, The 1895–1995, Congress Proceedings, vol. 2: Congress Lec-
formation, identification, and significance of DNA–protein tures, ICRR Society, 1995, pp. 400–403.
S. Barker et al. / Mutation Research 589 (2005) 111–135 135

[163] S.D. Desai, L.F. Liu, D. Vazquez-Abad, P. D’Arpa, Ubiquitin- trans-platinum complex showing higher antitumor activity
dependent destruction of topoisomerase I is stimulated by the than the cis congeners, J. Med. Chem. 36 (1993) 510–512.
antitumor drug camptothecin, J. Biol. Chem. 272 (1997) [167] M. Coluccia, A. Boccarelli, M.A. Mariggio, N. Cardellicchio,
24159–24164. P. Caputo, F.P. Intini, G. Natile, Platinum(II) complexes
[164] W.J. DeNeve, C.K. Everett, J.E. Suminski, F.A. Valeriote, containing iminoethers: a trans platinum antitumour agent,
Influence of WR2721 on DNA cross-linking by nitrogen Chem. Biol. Interact. 98 (1995) 251–266.
mustard in normal mouse bone marrow and leukemia cells [168] M. Coluccia, A. Nassi, A. Boccarelli, D. Giordano, N. Car-
in vivo, Cancer Res. 48 (1988) 6002–6005. dellicchio, F.P. Intini, G. Natile, A. Barletta, A. Paradiso, In
[165] J.Y. Wang, G. Prorok, W.P. Vaughan, Cytotoxicity, DNA vitro antitumour activity and cellular pharmacological proper-
cross-linking, and DNA single-strand breaks induced by ties of the platinum–iminoether complex trans-[PtCl2[E-
cyclophosphamide in a rat leukemia in vivo, Cancer Che- HN=C(OMe)Me]2], Int. J. Oncol. 15 (1999) 1039–1044.
mother. Pharmacol. 31 (1993) 381–386. [169] V. Orlando, H. Strutt, R. Paro, Analysis of chromatin structure
[166] M. Coluccia, A. Nassi, F. Loseto, A. Boccarelli, M.A. by in vivo formaldehyde cross-linking, Methods 11 (1997)
Mariggio, D. Giordano, F.P. Intini, P. Caputo, G. Natile, A 205–214.

Das könnte Ihnen auch gefallen