ArticlePDF AvailableLiterature Review

How do male germ cells handle DNA damage?

Authors:
  • Norwegian Institute for Air Research (NILU)

Abstract and Figures

Male reproductive health has received considerable attention in recent years. In addition to declining sperm quality, fertility problems and increased incidence of testicular cancer, there is accumulating evidence that genetic damage, in the form of unrepaired DNA lesions or de novo mutations, may be transmitted via sperm to the offspring. Such genetic damage may arise from environmental exposure or via endogenously formed reactive species, in stem cells or during spermatogenesis. Damaged testicular cells not removed by apoptosis rely on DNA repair for their genomic integrity to be preserved. To identify factors with potentially harmful effects on testicular cells and to characterise associated risk, a thorough understanding of repair mechanisms in these cells is of particular importance. Based on results from our own and other laboratories, we discuss the current knowledge of different pathways of excision repair in rodent and human testicular cells. It has become evident that, in human spermatogenic cells, some repair functions are indeed non-functional.
Limited repair of UV-C induced DNA lesion in rat male germ cells. To measure repair, the cells are treated with UV-C to introduce bulky DNA adducts such as cyclobutane pyrimidine dimers (CPDs) and incubated to allow repair in the absence (open bars) or presence (closed bars) of repair inhibitors arabinofuranoside C (AraC, 0.1 mM) plus hydroxyurea (HU, 2 mM) and analysed by the comet assay (A – D). Accumulation of DNA single-strand breaks (ssb) measured as tail moment indicates repair. The results show that crude cell suspensions (A) of rat testicular cells exhibit only limited repair following high doses of UV-C. Populations of different rat male germ cell types (spermatocytes (C) and round spermatids (D)) did not accumulate ssb following low doses of UV-C, while mononuclear blood cells (MNC) from rats as positive control cells (B) showed significant accumulation of ssb. NER in rat testicular cells was alternatively analysed with a modified comet method in which the bacterial T4endoV enzyme was used to cleave CPDs in DNA to quantify their presence in DNA in the form of ssb. Neither crude cell suspensions of rat testicular cells (E) nor isolated spermatocytes (F) removed significant numbers of CPDs within 24 h (open bars, no repair; closed bars, 24 h repair incubation), indicating very limited repair via NER in these cells. Furthermore, the results in panels (E) and (F) show that CPDs are indeed present since they are cleaved by T4endoV. In other experiments not shown here (Jansen et al., 2001), rat male germ cells were analysed while within their natural environment, the seminiferous tubule. Fragments of seminiferous tubules were exposed to the tissue-penetrating UV-B and incubated for repair within the tubules. The cells were subsequently squeezed out for comet analysis; results demonstrated a similar lack of incision as the results presented in panels A, C, and D. This also shows that lack of repair was not due to the cell isolation procedure or the absence of close contact with the nurturing and supporting Sertoli cells.
… 
Limited repair of some oxidative DNA lesions in human testicular cells. Repair of oxidative DNA bases is measured using the bacterial enzyme Formamidopyrimidine DNA N -glycosylase (Fpg) that is specific for the excision of oxidised pyrimidines such as 8-hydroxyguanine (8-oxoG). Oxidised base lesions are introduced into the cellular DNA using the photo-activated oxidising agent Ro 12-9786 (6 A M) plus cold visible light (5 min). The number of Fpg- sensitive DNA lesions is measured as single-strand breaks (ssb) in a modified alkaline elution assay. The damage levels are calculated from elution profiles and are expressed as F normalised area above curve _ (NAAC) (Brunborg et al., 1996). In these assays, genomic DNA is treated with an excess amount of Fpg crude extract to transform Fpg-sensitive DNA lesions into ssb. (A) Repair of Fpg-sensitive DNA lesions by normal primary human fibroblasts. Cell samples were treated with Ro 12-9786 (6 A M) plus light (5 min) to introduce oxidative base damage and treated with (diamonds) or without Fpg (triangles), respectively, after being allowed to repair. Solid and broken lines show their mean values. Circles and squares represent control cell samples treated with or without Fpg, respectively. (B – C) Repair of Fpg-sensitive DNA lesions by human testicular cells (B) from one individual testis biopsy and rat testicular cells (C, 3 A M Ro 12-9786). For symbols, see panel A and box in panel A. The results demonstrate that human testicular cells exhibit only limited repair of Fpg-sensitive DNA lesions as opposed to rapid repair by both rat testicular cells and human fibroblasts. On the other hand, we have observed that the repair of other oxidised DNA lesions, such as some oxidised purines, is efficient in both human and rat testicular cells (data not shown here; Olsen et al., 2003).
… 
Content may be subject to copyright.
Review
How do male germ cells handle DNA damage?
Ann-Karin Olsen, Birgitte Lindeman, Richard Wiger, Nur Duale, Gunnar Brunborg*
Norwegian Institute of Public Health, Department of Chemical Toxicology, P.O. Box 4404 Nydalen, N-0403 Oslo, Norway
Received 12 January 2005; revised 13 January 2005; accepted 21 January 2005
Available online 27 July 2005
Abstract
Male reproductive health has received considerable attention in recent years. In addition to declining sperm quality, fertility problems and
increased incidence of testicular cancer, there is accumulating evidence that genetic damage, in the form of unrepaired DNA lesions or de
novo mutations, may be transmitted via sperm to the offspring. Such genetic damage may arise from environmental exposure or via
endogenously formed reactive species, in stem cells or during spermatogenesis. Damaged testicular cells not removed by apoptosis rely on
DNA repair for their genomic integrity to be preserved. To identify factors with potentially harmful effects on testicular cells and to
characterise associated risk, a thorough understanding of repair mechanisms in these cells is of particular importance. Based on results from
our own and other laboratories, we discuss the current knowledge of different pathways of excision repair in rodent and human testicular
cells. It has become evident that, in human spermatogenic cells, some repair functions are indeed non-functional.
D2005 Elsevier Inc. All rights reserved.
Keywords: Testis; DNA repair; Male germ cells; Sperm; Chemical toxicity
Contents
DNA damage, fertility and reproduction ............................................. S521
Spermatogenesis ......................................................... S522
DNA repair............................................................ S522
DNA excision repair in germ cells ................................................ S522
Nucleotide Excision Repair (NER)................................................ S524
Base Excision Repair (BER) ................................................... S527
What happens with the germ cells containing unrepaired DNA lesions? ............................ S529
Concluding remarks ....................................................... S529
Epilogue ............................................................. S530
Acknowledgments ........................................................ S530
References ............................................................ S530
DNA damage, fertility and reproduction
Evidence from both animal studies and from human
epidemiology is now accumulating, indicating that environ-
mental agents can interact with the male genome, causing
specific types of genetic change in sperm that affect fertility.
Human spermatozoa frequently possess high levels of
nuclear DNA damage (Sun et al., 1997; Irvine et al.,
2000; Evenson et al., 2002). Even with significant levels of
DNA damage, sperm retain the ability to fertilise oocytes;
however, the subsequent development may be altered or
lead to zygote arrest. Further implications are early
0041-008X/$ - see front matter D2005 Elsevier Inc. All rights reserved.
doi:10.1016/j.taap.2005.01.060
* Corresponding author. Fax: +47 2204 2686.
E-mail address: gunnar.brunborg@fhi.no (G. Brunborg).
Toxicology and Applied Pharmacology 207 (2005) S521 – S531
www.elsevier.com/locate/ytaap
abortions, congenital malformations and disease including
cancer in the offspring (Ahmadi and Ng, 1999; Zenzes et al.,
1999b; Brinkworth, 2000). The importance of DNA damage
is further illustrated by experiences with assisted reproduc-
tion (Zitzmann et al., 2003; Sun et al., 1997; Twigg et al.,
1998; Morris et al., 2002) as recently reviewed by Sharma et
al., 2004.
In addition to the data discussed by Diana Anderson in
this symposium, there are other cases of paternally trans-
mitted malformations via mechanisms not yet understood. A
cluster of malformations was discovered (e.g. club foot)
among children of fathers previously working on a marine
vessel in Norway (RR = 4.6, CI 2.6 8.4; Professor Bente E.
Moen, personal communication). The cause of this apparent
increased risk is however not known.
The genetic constitution of the offspring depends
logically on the integrity of the sperm and the egg DNA,
both expected to be of equal importance. Traditionally,
concern about possible developmental toxic effects from
environmental agents has focussed on the maternal con-
tribution, whereas paternal aspects have been generally
ignored. Correspondingly, chemicals have been examined
for developmental toxicity using maternal and not paternal
test systems.
Among environmental factors known to mediate DNA
damage in sperm, smoking is well-documented. Heavy
smokers exhibit increased levels of strand breaks in their
spermatozoa (Fraga et al., 1996; Potts et al., 1999), and it
has now been shown that smoking causes more abnormal
sperm, increased levels of oxidative DNA lesions such as
7,8-dihydro-8-oxoguanine (8-oxoG) and reduced fecundity
(Zitzmann et al., 2003). Furthermore, children of smoking
fathers may have an increased risk of developing childhood
cancer (Ji et al., 1997).
Spermatogenesis
Spermatogenesis is the process during which sperm is
produced from progenitor spermatogonia and is initiated
when the male has attained sexual maturity. It comprises
three main phases: the spermatogonial stage, the meiotic
stage and the spermiogenesis stage; in general, the
duration of each of these stages is approximately equal.
Some more details are given in Fig. 1 and its legend; the
reader is otherwise referred to recent reviews (Holstein et
al., 2003).
Gametes may be susceptible to environmental genoto-
xicants at different stages of spermatogenesis. Unrepaired
DNA lesions in spermatogonia may be fixed as mutations
during replication giving rise to clones of mutated sperm.
Cells in later spermatogenic stages do not replicate, and
fixation of premutagenic DNA lesions as mutations depends
on replication-independent DNA synthesis associated with
recombination or DNA repair. Alternatively, mutations may
be fixed during repair in the zygote.
DNA repair
There are several lines of defense against the induction
and persistence of DNA damage in cells. Firstly, there
are agents that prevent the formation of DNA damage,
such as detoxifying peptides and proteins, and oxyradical
scavengers, such as vitamins E and C. Secondly, DNA
damage already present in the genome may be sensed by
DNA damage response pathways, and the damage may
be removed by an array of DNA repair pathways to
reduce the possibility for inducing mutations. Thirdly, the
damaged cell itself may be eliminated by spontaneous
death or apoptosis.
DNA damage can have a variety of biological
ramifications including inhibition of transcription and/or
replication, ultimately causing cell death, mutations and
cancer. Constant excision and replacement of damaged
nucleotide residues by DNA repair are required to
counteract the occurrence of such events. The manner
in which a DNA lesion is removed depends on the nature
of the DNA lesion itself; for one specific type of DNA
lesion, there is often a redundancy of pathways that
function as back-ups for each other. Most of the repair
mechanisms are not absolutely exact and error-free,
whereas commonly occurring DNA lesions clearly must
be removed with high accuracy.
In this report, the focus is on excision repair, in which
an altered nucleotide residue is replaced following
removal of a segment of the damaged strand followed
by synthesis of intact DNA (Fig. 2). One pathway,
Nucleotide Excision Repair (NER), removes predomi-
nantly helix-distorting lesions that mostly originate via
environmental agents (Fig. 2). A second pathway, Base
Excision Repair (BER), removes mainly common mod-
ifications such as oxidations, methylations, deaminations
and base losses (Fig. 3). These types of lesions are
caused by both exogenous agents and endogenous
processes.
DNA damage can also be repaired by other mechanisms
not discussed here such as direct reversal, mismatch repair
(MMR), homologous recombination (HR) and non-homol-
ogous end joining (NHEJ) (recently reviewed in germ cells
by Baarends et al., 2001).
DNA excision repair in germ cells
There is limited information on repair functions in
male germ cells. In early studies, this was probably
related to a lack of appropriate methodologies, a situation
which has improved during recent years. Traditionally,
DNA repair in male germ cells has mostly been studied
using assays such as unscheduled DNA synthesis (UDS),
alkaline elution, the comet assay, and immunochemical
methods. Furthermore, the testes represent a challenge
with respect to the isolation and cultivation of the
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521– S531S522
different male germ cell types. DNA single-strand breaks
(SSB) in male germ cells of the earlier stages of
spermatogenesis from both hamsters and rats are repaired
as efficiently as in several somatic cell types, whereas
elongated spermatids (see Fig. 1) exhibited no repair (van
Loon et al., 1991, 1993). More recently, our laboratory
Fig. 1. A schematic presentation of spermatogenesis. The general organisation of spermatogenesis is essentially the same in all mammals and can be divided
into three phases, the spermatogonial, the meiotic and the spermiogenesis stages. The spermatogonial or proliferative stage starts with a division of
spermatogonial stem cells (originating from primordial germ cells during development) into two daughter cells, one of which enters the process of
spermatogenesis, while the other remains as a stem cell. This is the period of active replicative DNA synthesis producing different types of spermatogonia. The
number of cell divisions during this stage varies with species, but ultimately type B spermatogonia give rise to tetraploid primary spermatocytes in the
preleptotene stage. During the first part (prophase) of the meiotic stage, genetic recombination takes place after which a first reduction division gives rise to
secondary 2n spermatocytes, and subsequently the second reduction division results in haploid round spermatids. During spermiogenesis, extensive changes
occur in spermatids, including nuclear condensation, leading to spermatozoa. Some additional details and the duration of the stages are indicated. Further
details may be found in the comprehensive review by Holstein et al. (2003).
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521 S531 S523
made similar observations with human (Olsen et al.,
2003), mouse and rat (Wiger et al., unpublished) male
germ cells.
Nucleotide Excision Repair (NER)
NER is a major DNA repair pathway that acts on a wide
variety of helix-distorting DNA lesions (bulky adducts) such
as those caused by UV light (cyclobutane pyrimidine dimer,
(CPD) and 6-4 photoproduct (6-4PP)) and by a range of
exogenous chemicals including environmental agents such as
benzo(a)pyrene (B[a]P) and aflatoxin B
1
and chemother-
apeutic agents such as cisplatin. A defect in one of the repair
proteins results in recessive syndromes such as xeroderma
pigmentosum (XP), associated with very high (1000-fold
increased) cancer risk (Cleaver, 1989).
NER involves recognition of the DNA lesion, incision of
the DNA strand containing the lesion, followed by DNA
synthesis and ligation to replace an excised oligonucleotide
(there are several excellent reviews including Wood, 1997;
de Laat et al., 1999; Mitchell et al., 2003). There are two
NER subpathways (Fig. 2), the global genomic repair
(GGR) that repairs DNA lesions in the entire genome and
the transcription-coupled repair (TCR) that removes DNA
lesions that block RNA synthesis, i.e. in transcribed genes
(Tornaletti and Hanawalt, 1999).
In total, 25 or more proteins are involved in NER (Fig. 2,
abbreviations for each enzyme are given in the figure
legend). GGR is initiated by binding of XPC-hHR23B to
Fig. 2. An outline of the Nucleotide Excision Repair (NER) pathway. See text for description of the pathway. Abbreviations in chronological order: XPA–G,
xeroderma pigmentosum complementation group AG; hHR23B, human homologue of yeast RAD23B; RNA Pol II, RNA polymerase II; CSA and CSB,
Cockayne syndrome factors A and B; TFIIH, general transcription factor IIH; ERCC1, excision repair cross complementing group 1 protein; RPA, replication
protein A; PCNA, proliferating cell nuclear antigen; RFC, replication factor C; Poly/E, DNA polymerase delta/epsilon; Lig1, DNA ligase 1.
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521– S531S524
disrupted base pairs. During TCR, lesions that block the
RNA polymerase are detected, and the polymerase is
displaced, making the DNA lesion accessible for repair;
this requires at least two TCR-specific factors: CSA and
CSB. During both GGR and TCR, XPG associates and the
general transcription factor TFIIH (a multi-protein complex
that includes the two helicases XPB and XPD) unwinds
about 30 base pairs surrounding the DNA lesion. The
subsequent steps of GGR and TCR are believed to be
identical. XPA associates with the single-strand binding
protein RPA to form a complex that confirms the presence
of DNA damage by sensing a simultaneous backbone and
base pair distortion of DNA (Missura et al., 2001). The
ERCC1-XPF endonuclease complex and the XPG endonu-
clease cleave the damaged strand 5Vand 3Vrelative to the
DNA lesion, generating a 24- to 32-base oligonucleotide
containing the lesion. New DNA is synthesised, and the gap
is ligated.
A third and interesting subpathway of NER has recently
been proposed and given the name differentiation-associated
repair (DAR). As opposed to TCR, in which repair is rapid in
the transcribed strand of expressed genes, during DAR in
terminally differentiated cells, both the non-transcribed and
the transcribed strands are efficiently repaired, with DAR
taking care of the non-transcribed strand which is otherwise
repaired by GGR in undifferentiated cells (Nouspikel and
Hanawalt, 2002). The significance and general existence of
DAR are however still debated.
Fig. 3. An outline of the Base Excision Pathway (BER) pathway. See text for description of the pathway. Abbreviations in chronological order: HAP1, human AP-
endonuclease 1; Polh, DNA polymerase h; XRCC1, X-ray cross complementing protein 1; LigIII, DNA ligase III; PCNA, proliferating cell nuclear antigen; RFC,
replication factor C; Poly(, DNA polymerase y(; FEN1, Flap endonuclease; Lig1, DNA ligase 1. The figure is modified from Ide and Kotera (2004).
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521 S531 S525
The efficiency of repair of different kinds of lesions varies
by several orders of magnitude. The rate of repair for GGR is
strongly dependent on the type of lesion, i.e. 6-4PPs are
removed much faster from the genome than CPDs (¨1h
compared to ¨6 h in Chinese hamster ovary cells, respec-
tively; Nairn et al., 1989).
About 10 years ago, we observed indications of very
low repair in rat testicular cells exposed to UV-C
(Brunborg et al., 1995). This was based on a lack of
accumulated DNA single-strand breaks (SSBs) in the
presence of repair inhibitors, using the alkaline elution
assay or the alkaline comet assay. The rationale behind
Fig. 4. Limited repair of UV-C induced DNA lesion in rat male germ cells. To measure repair, the cells are treated with UV-C to introduce bulky DNA adducts
such as cyclobutane pyrimidine dimers (CPDs) and incubated to allow repair in the absence (open bars) or presence (closed bars) of repair inhibitors
arabinofuranoside C (AraC, 0.1 mM) plus hydroxyurea (HU, 2 mM) and analysed by the comet assay (A D). Accumulation of DNA single-strand breaks
(ssb) measured as tail moment indicates repair. The results show that crude cell suspensions (A) of rat testicular cells exhibit only limited repair following high
doses of UV-C. Populations of different rat male germ cell types (spermatocytes (C) and round spermatids (D)) did not accumulate ssb following low doses of
UV-C, while mononuclear blood cells (MNC) from rats as positive control cells (B) showed significant accumulation of ssb. NER in rat testicular cells was
alternatively analysed with a modified comet method in which the bacterial T4endoV enzyme was used to cleave CPDs in DNA to quantify their presence in
DNA in the form of ssb. Neither crude cell suspensions of rat testicular cells (E) nor isolated spermatocytes (F) removed significant numbers of CPDs within
24 h (open bars, no repair; closed bars, 24 h repair incubation), indicating very limited repair via NER in these cells. Furthermore, the results in panels (E) and
(F) show that CPDs are indeed present since they are cleaved by T4endoV. In other experiments not shown here (Jansen et al., 2001), rat male germ cells were
analysed while within their natural environment, the seminiferous tubule. Fragments of seminiferous tubules were exposed to the tissue-penetrating UV-B and
incubated for repair within the tubules. The cells were subsequently squeezed out for comet analysis; results demonstrated a similar lack of incision as the
results presented in panels A, C, and D. This also shows that lack of repair was not due to the cell isolation procedure or the absence of close contact with the
nurturing and supporting Sertoli cells.
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521– S531S526
using repair inhibitors is as follows: UV-C induces bulky
lesions repaired by NER; such lesions are normally
recognised and incised by the cells’ own repair machi-
nery and are thus converted into strand breaks. These
breaks stay open and accumulate when cells are
incubated to allow repair in a medium containing repair
inhibitors (blocking DNA polymerase and deoxyribonu-
cleotide synthesis; see legend to Fig. 4). This is a
sensitive way of measuring excision repair activities in
live cells. We have subsequently studied NER in rat,
mouse and human testicular cells to get a better under-
standing of how environmental agents that cause bulky
lesions in DNA are handled in these cells. Removal of
CPDs was poor in rat testicular cells in suspension (Fig.
4), as well as in cells in their normal environment within
intact seminiferous tubules, using low doses of UV. In
enriched cell samples of rat spermatocytes of zygotene/
leptotene, mid-pachytene and diplotene stages exposed to
30 J/m
2
UV-C, TCR was also low (measured in the
transcribed strand of the meiosis specific gene SCP1)
(Jansen et al., 2001). Incision activities of cell extracts
were measured using an oligonucleotide containing a
defined lesion; excision was high with early- and mid-
pachytene spermatocytes as opposed to extracts from
other stages (diplotene spermatocytes and round sperma-
tids). It is not unexpected that protein extracts exhibit
proficient repair activities (as recently also shown by
Gospodinov et al., 2003), while in the live cell repair
may still be absent since it depends on the distribution,
spatial translocation and association of repair proteins. We
hypothesised that NER proteins are sequestered by
mispaired regions in DNA involved in synapsis and
recombination and that this explains the lack of NER
activity in premeiotic cells (Jansen et al., 2001). Later
experiments have indicated that some NER-associated
proteins may be expressed in low amounts in human
male germ cells (unpublished). At the level of mRNA,
however, transcripts of most NER-related enzymes are
expressed in high amounts (Li et al., 1996; van der Spek
et al., 1996; Cheng et al., 1999; Shannon et al., 1999).
More recently, Nouspikel and Hanawalt (2002) have
discussed observations that NER activities are highly differ-
ent among various cell types; GGR is generally switched off
in terminally differentiated cells and is replaced by DAR. In
this context, our results with rat male germ cells (Jansen et al.,
2001) are paradoxical since they apparently exhibit no TCR
(and hence no DAR).
Low NER is in compliance with the observations
recently reviewed by Sotomayor and Sega (2000).In
studies of UDS in male germ cells exposed to different
agents, as many as 59 chemicals plus UV and X-rays
were tested in spermatogenic cells of humans, rabbits,
rats and mice. Although these aspects were not discussed
in the review, in general, agents inducing DNA lesions that
are removable by NER did not show UDS (2-AAF,
aflatoxin B
1
,B[a]P, N-OH-AAF).
Base Excision Repair (BER)
BER is initiated with the release of altered bases by DNA
glycosylases (Fig. 3), as first reported by Thomas Lindahl
30 years ago (Lindahl, 1974). DNA glycosylases recognise
and excise aberrant bases that cause minor structural
changes in DNA, and each DNA glycosylase recognises a
specific set of DNA substrates. The excision of the base
generates apurinic/apyrimidinic sites (AP sites) followed by
endonuclease cleavage, re-synthesis and ligation. Some
DNA glycosylases are bi-functional, that is, besides their
DNA N-glycosylase activity, they have the ability to cleave
at AP sites (AP-lyase). At least eleven human DNA
glycosylases are known to date. BER is subdivided into
short-patch repair (SPR, 1 nucleotide (nt)) and long-patch
repair (LPR, 2 10 nt) (Fig. 3). During SPR, the major AP
endonuclease HAP1, or the 5V-terminal deoxyribose phos-
phatase (5V-dRPase) activity of Polh, trims the DNA ends
into suitable substrates for new DNA synthesis. Polh
incorporates one nucleotide, and the nick is sealed by
DNA ligase III/XRCC1. During LPR, Polyor (incorporates
several nucleotides in association with RPA and PCNA,
generating a 5V-dRP flap structure. This structure is nicked
by FEN1, and the gap is sealed by DNA ligase I. Similar to
NER, DNA lesions that inhibit or block transcription, such
as 8-oxoG, are probably removed by BER in a transcription-
coupled manner (Leadon and Cooper, 1993; Cooper et al.,
1997; Le Page et al., 2000), partly using the same enzymes
as in NER-TCR.
Many altered bases are repaired by BER; we have
studied some of them and their associated DNA glycosy-
lases and downstream activities. Uracil residues may lead to
mutations and occur in the genome either as a result of
erroneous incorporation via replication or by deamination of
thymine. We found that the most important uracil-DNA
glycosylase (UDG) was present in human and rat testicular
cells in amounts equal to somatic tissues (Olsen et al.,
2001). This is based upon measuring enzymatic activities
and by immunochemical detection of the enzyme. UDG has
one nuclear and one mitochondrial isoform, and both were
detected in our analyses. The data indicate that uracil is
removed from the male germ cell genome; this is
corroborated by Intano et al. (2001), who reported UDG
also in mouse testicular cells. Furthermore, Grippo et al.
(1982) detected human uracil-DNA glycosylase (UNG)
activity in the DNA-synthesising male germ cells. The
presence of at least five different and partly overlapping
DNA glycosylases for removal of uracil suggests that this
repair is important. In cases of diets low in folic acid, uracil
is incorporated into the genome in higher amounts, and
therefore pregnant women are recommended to add folic
acid to their diet to prevent the embryo from developing
spina bifida. Mice deficient in UDG have been constructed
and exhibit a 50% decreased ability to excise uracil (Nilsen
et al., 2000). These mice retain the ability to produce
offspring, indicating that UDG is not essential for repro-
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521 S531 S527
duction. The adverse effects detected in these mice include
impaired immunoglobulin production (Imai et al., 2003),
higher morbidity and a higher chance of developing B-cell
lymphomas (Nilsen et al., 2003). It has been suggested that
the reason for this apparently mild phenotype is the
redundancy of other uracil-DNA glycosylases, such as
SMUG (single-strand selective monofunctional uracil-
DNA glycosylase), that act as back-ups. Alkylations
represent a major class of DNA lesions; several types are
repaired via the methylpurine DNA glycosylase (MPG).
Mpg mRNA is abundant in mice testes (Engelward et al.,
1993; Kim et al., 2000). Other proteins that repair
alkylations are the direct acting ‘‘suicide protein’’ O
6
-
Methylguanine-DNA-glycosylase (MGMT) and the recently
identified AlkB (alkylation repair) human homologue
(ABH) proteins (Duncan et al., 2002; Aas et al., 2003),
each with their specific DNA substrates. We found that the
MGMT protein was expressed in normal amounts in human
and rat male germ cells (unpublished), similar to other
investigators’ findings in mice (Thompson et al., 2000). In
our studies, MPG is present in greater amounts in human
and rat male germ cells compared to somatic cells (Olsen et
al., 2001). Minor differences were observed between
different cellular stages of rat spermatogenesis and spermio-
genesis. DNA lesions induced by exposure to the mutagen
methyl-methane sulfonate (MMS) were 5-fold higher in rat
compared to human male germ cells, indicating major
differences in sensitivity between the two species. Repair of
methylated DNA studied at the cellular level was efficient in
both human and rat male germ cells, in primary spermato-
cytes as well as round spermatids, compared to primary
somatic cell types. This proficient repair is consistent with
the lack of increased mutation rates in male germ cells of
Big Blue transgenic mice treated with MMS (Ashby et al.,
1997); large insertions and deletions are however not scored
in the latter assay. Hence, male germ cells seem well
protected against DNA methylations such as those inflicted
by MMS; such lesions arise spontaneously or via the
environment. Other mouse germ cell alkylating mutagens,
such as ethyl nitrosourea (ENU) and isopropyl methane-
sulphonate (iPMS), attack other base positions and do
induce mutations (Ashby et al., 1997). Alkylating agents
also attack other components of a cell besides DNA bases,
and it is now believed that some of these disturb chromatin
packaging during spermiogenesis.
Oxidised lesions and their repair in male germ cells are of
special interest. One prevalent and important lesion is 8-
oxoG arising from oxidised guanine or misincorporated
oxidised dGTP. 8-oxoG is mutagenic and is an inevitable
consequence of oxidative metabolism but is also induced by
many environmental mutagens including cigarette smoke
and exhaust (Shen et al., 1997). Other important lesions
include oxidised pyrimidines such as thymine glycols (TG)
and 5-hydroxycytosine (5-OHC). We have studied the repair
Fig. 5. Limited repair of some oxidative DNA lesions in human testicular cells. Repair of oxidative DNA bases is measured using the bacterial enzyme
Formamidopyrimidine DNA N-glycosylase (Fpg) that is specific for the excision of oxidised pyrimidines such as 8-hydroxyguanine (8-oxoG). Oxidised base
lesions are introduced into the cellular DNA using the photo-activated oxidising agent Ro 12-9786 (6 AM) plus cold visible light (5 min). The number of Fpg-
sensitive DNA lesions is measured as single-strand breaks (ssb) in a modified alkaline elution assay. The damage levels are calculated from elution profiles and
are expressed as Fnormalised area above curve_(NAAC) (Brunborg et al., 1996). In these assays, genomic DNA is treated with an excess amount of Fpg crude
extract to transform Fpg-sensitive DNA lesions into ssb. (A) Repair of Fpg-sensitive DNA lesions by normal primary human fibroblasts. Cell samples were
treated with Ro 12-9786 (6 AM) plus light (5 min) to introduce oxidative base damage and treated with (diamonds) or without Fpg (triangles), respectively, after
being allowed to repair. Solid and broken lines show their mean values. Circles and squares represent control cell samples treated with or without Fpg,
respectively. (B– C) Repair of Fpg-sensitive DNA lesions by human testicular cells (B) from one individual testis biopsy and rat testicular cells (C, 3 AMRo
12-9786). For symbols, see panel A and box in panel A. The results demonstrate that human testicular cells exhibit only limited repair of Fpg-sensitive DNA
lesions as opposed to rapid repair by both rat testicular cells and human fibroblasts. On the other hand, we have observed that the repair of other oxidised DNA
lesions, such as some oxidised purines, is efficient in both human and rat testicular cells (data not shown here; Olsen et al., 2003).
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521– S531S528
of these lesions in rodent and human male germ cells using
cellular extracts, identifying the relevant proteins and by
measuring active repair in live cells (Fig. 5). Oxidised
purines such as 8-oxoG were efficiently repaired in rat (Olsen
et al., 2003) and mouse (unpublished) spermatogenic cells.
However, human testicular cells from several individuals
showed no repair of these lesions, a result which we found to
be most unexpected. On the other hand, oxidised pyrimidines
such as TG were efficiently repaired in both the human and
the rat. In addition to the cellular assays, enzymatic activities
and protein levels of the relevant repair enzymes were
measured and were in accordance with the efficiency of
cellular repair. A very high level of Ogg1 mRNA was
reported in mouse testis (Rosenquist et al., 1997), whereas in
human tissues including the testis, OGG1 mRNA is
ubiquitously expressed (Radicella et al., 1997; Nishioka et
al., 1999). In conclusion, it appears that human male germ
cells may be particularly sensitive to DNA oxidation.
Mice and rats are by far the preferred species for
mutagenicity testing and reproductive toxicity studies. Since
it has been shown that humans may be different with respect
to the repair of important pre-mutagenic lesions, the results
from toxicological tests using laboratory species should be
interpreted with great care. A transgenic model has been
constructed in which the main DNA glycosylase (mOGG1)
for removal of 8-oxoG has been knocked out (Klungland et
al., 1999). Nuclear testicular extracts showed no incision of
8-oxoG (Klungland et al., 1999), also suggesting that the
testis has little back-up activity; indeed, we detected no
cellular repair of oxidative lesions in the male germ cells
from these mice (unpublished observations). We are
currently evaluating the mOGG1 KO mouse as a model
for human male germ cell toxicity testing. BER-related
enzymes involved in the later stages of the pathway seem to
be present at levels sufficient for functional repair. A number
of studies support this view. Nuclear extracts from mouse
spermatogenic germ cells contain high amounts of Ligase I,
Ape (mouse homologue of the human HAP1), Ligase III,
Xrcc1 and Polh, compared to levels in somatic cells (Intano
et al., 2001). Polhis highly expressed in bovine pachytene
spermatocytes (Hirose et al., 1989). Both Polhand DNA
ligase I are present in the rat (Prasad et al., 1996). DNA
ligase activities in the mouse were higher in premeiotic
spermatogonia and spermatocytes than in liver and bone
marrow cells (Higashitani et al., 1990). High levels of DNA
ligases I and III were also found in bovine testis (Husain et
al., 1995). Human DNA ligase III is ubiquitously expressed
at low levels, except in the testis, where the steady state
levels of DNA ligase III mRNA are at least 10-fold higher
than those detected in other tissues and cells (Chen et al.,
1995). The DNA ligase I mRNA expression in the testes
correlated with the proportion of proliferative spermatogo-
nia, in agreement with the previously defined role of this
enzyme in DNA replication. In contrast, elevated levels of
DNA ligase III mRNA were observed in primary sperma-
tocytes undergoing recombination prior to the first meiotic
division. This suggests that DNA ligase III seals DNA strand
breaks that arise during the process of meiotic recombination
in germ cells (Chen et al., 1995).
What happens with the germ cells containing unrepaired
DNA lesions?
The poor removal of bulky DNA adducts and oxidised
purines in human testicular cells are reflected in an
accumulation of DNA damage such as 8-oxoG and
benzo(a)pyrene adducts in human sperm (Sun et al., 1997;
Irvine et al., 2000; Evenson et al., 2002; Zenzes et al.,
1999b). These lesions are clearly environmentally related,
with smoking as a good (or bad) example (Shen et al., 1997;
Zenzes et al., 1999b).
Apoptosis then represents an additional mechanism for
removing cells with DNA lesions. Spontaneous apoptosis
does occur in all cell types of the testis during normal
spermatogenesis in man (Oldereid et al., 2001), and
testicular cells are very sensitive to apoptotic stimuli such
as high-dose chemotherapy (Spierings et al., 2003). It is
likely that the total level of DNA lesions is of decisive
importance for the fate of the spermatogenic cell.
On the other hand, there is evidence that B(a)P adducts
can indeed accumulate without eliciting apoptosis in
spermatogenic cells since ejaculated sperm from smokers
contain more B(a)P adducts than from non-smokers (Zenzes
et al., 1999a). Furthermore, such adducts derived from the
father do not prevent fertilisation and are persistent even at
the blastocyst level (Zenzes et al., 1999b), also indicating
that they are not always completely repaired in the fertilised
oocyte. A recent publication (Zitzmann et al., 2003) showed
that assisted fertilisation was 40% less likely to produce
a live offspring if the father was a smoker, compared to
non-smokers.
Concluding remarks
A question that emerges from these results is whether the
shutting off of some repair functions has a purpose. Several
possibilities may be envisaged. One might speculate that,
along with other functions in spermatogenic cells, energy
consumption is minimised (with priority directed to
spermatogenesis-associated processes such as meiotic
recombination and repackaging of DNA). A further
possibility is the need for a finely tuned balance (controlled
by the presence of some error-prone repair) between
preservation of the genetic integrity vs. the need for some
de novo germ-line mutations to drive evolution. Thirdly, the
presence of DNA lesions in a sperm that has reached the
oocyte may serve as a direct indicator for the oocyte as to
the genotoxic load that the sperm has been subjected to; the
oocyte may then either attempt repair or terminate the
process.
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521 S531 S529
Epilogue
In the event that sperm cell DNA has not been repaired and
apoptosis was not initiated, then – like all other problems
we leave it up to the mother to solve the problem!
Acknowledgments
We thank Professor Bente E. Moen (Department of
Public Health and Primary Health Care, University of
Bergen, Norway) for the permission to use unpublished
data on malformations among naval serviceman. This
work has been supported by the European Union (ENV4-
CT95-0204 and QLK4-CT-2002-02198), and by The
Norwegian Research Council (Grants Nos 129614/310
and 148703/310).
References
Aas, P.A., Otterlei, M., Falnes, P.O., Vagbo, C.B., Skorpen, F., Akbari, M.,
Sundheim, O., Bjoras, M., Slupphaug, G., Seeberg, E., Krokan, H.E.,
2003. Human and bacterial oxidative demethylases repair alkylation
damage in both RNA and DNA. Nature 421, 859 863.
Ahmadi, A., Ng, S.C., 1999. Fertilizing ability of DNA-damaged
spermatozoa. J. Exp. Zool. 284, 696– 704.
Ashby, J., Gorelick, N.J., Shelby, M.D., 1997. Mutation assays in male
germ cells from transgenic mice: overview of study and conclusions.
Mutat. Res. 388, 111 – 122.
Baarends, W.M., van der, L.R., Grootegoed, J.A., 2001. DNA repair
mechanisms and gametogenesis. Reproduction 121, 3139.
Brinkworth, M.H., 2000. Paternal transmission of genetic damage: findings
in animals and humans. Int. J. Androl. 23, 123 – 135.
Brunborg, G., Holme, J.A., Hongslo, J.K., 1995. Inhibitory effects of
paracetamol on DNA repair in mammalian cells. Mutat. Res. 342,
157 – 170.
Brunborg, G., Soderlund, E.J., Holme, J.A., Dybing, E., 1996. Organ-
specific and transplacental DNA damage and its repair in rats treated
with 1,2-dibromo-3-chloropropane. Chem.-Biol. Interact. 101, 33– 48.
Chen, J., Tomkinson, A.E., Ramos, W., Mackey, Z.B., Danehower, S.,
Walter, C.A., Schultz, R.A., Besterman, J.M., Husain, I., 1995.
Mammalian DNA ligase III: molecular cloning, chromosomal local-
ization, and expression in spermatocytes undergoing meiotic recombi-
nation. Mol. Cell. Biol. 15, 5412– 5422.
Cheng, L., Guan, Y., Li, L., Legerski, R.J., Einspahr, J., Bangert, J., Alberts,
D.S., Wei, Q., 1999. Expression in normal human tissues of five
nucleotide excision repair genes measured simultaneously by multiplex
reverse transcription-polymerase chain reaction. Cancer Epidemiol.
Biomark. Prev. 8, 801 – 807.
Cleaver, J.E., 1989. DNA repair in man. Birth Defects Orig. Artic. Ser. 25,
61 – 82.
Cooper, P.K., Nouspikel, T., Clarkson, S.G., Leadon, S.A., 1997. Defective
transcription-coupled repair of oxidative base damage in Cockayne
syndrome patients from XP group G. Science 275, 990 – 993.
de Laat, W.L., Jaspers, N.G., Hoeijmakers, J.H., 1999. Molecular
mechanism of nucleotide excision repair. Genes Dev. 13, 768 785.
Duncan, T., Trewick, S.C., Koivisto, P., Bates, P.A., Lindahl, T., Sedgwick,
B., 2002. Reversal of DNA alkylation damage by two human
dioxygenases. Proc. Natl. Acad. Sci. U.S.A. 99, 16660 – 16665.
Engelward, B.P., Boosalis, M.S., Chen, B.J., Deng, Z., Siciliano, M.J.,
Samson, L.D., 1993. Cloning and characterization of a mouse 3-
methyladenine/7-methyl-guanine/3-methylguanine DNA glycosylase
cDNA whose gene maps to chromosome 11. Carcinogenesis 14,
175 – 181.
Evenson, D.P., Larson, K.L., Jost, L.K., 2002. Sperm chromatin structure
assay: its clinical use for detecting sperm DNA fragmentation in male
infertility and comparisons with other techniques. J. Androl. 23, 25 – 43.
Fraga, C.G., Motchnik, P.A., Wyrobek, A.J., Rempel, D.M., Ames, B.N.,
1996. Smoking and low antioxidant levels increase oxidative damage to
sperm DNA. Mutat. Res. 351, 199– 203.
Gospodinov, A., Ivanov, R., Anachkova, B., Russev, G., 2003. Nucleotide
excision repair rates in rat tissues. Eur. J. Biochem. 270, 1000– 1005.
Grippo, P., Orlando, P., Lococrondo, G., Geremia, R., 1982. Uracil-DNA
glycosylase in meiotic and post meiotic male germ cells of the mouse.
Prog. Clin. Biol. Res. 85, 389 396.
Higashitani, A., Tabata, S., Endo, H., Hotta, Y., 1990. Purification of DNA
ligases from mouse testis and their behavior during meiosis. Cell Struct.
Funct. 15, 67 72.
Hirose, F., Hotta, Y., Yamaguchi, M., Matsukage, A., 1989. Difference in
the expression level of DNA polymerase beta among mouse tissues:
high expression in the pachytene spermatocyte. Exp. Cell Res. 181,
169 – 180.
Holstein, A.F., Schulze, W., Davidoff, M., 2003. Understanding spermato-
genesis is a prerequisite for treatment. Reprod. Biol. Endocrinol. 1, 107.
Husain, I., Tomkinson, A.E., Burkhart, W.A., Moyer, M.B., Ramos, W.,
Mackey, Z.B., Besterman, J.M., Chen, J., 1995. Purification and
characterization of DNA ligase III from bovine testes. Homology
with DNA ligase II and vaccinia DNA ligase. J. Biol. Chem. 270,
9683 – 9690.
Ide, H., Kotera, M., 2004. Human DNA glycosylases involved in the repair
of oxidatively damaged DNA. Biol. Pharm. Bull. 27, 480 – 485.
Imai, K., Slupphaug, G., Lee, W.I., Revy, P., Nonoyama, S., Catalan, N.,
Yel, L., Forveille, M., Kavli, B., Krokan, H.E., Ochs, H.D., Fischer, A.,
Durandy, A., 2003. Human uracil-DNA glycosylase deficiency asso-
ciated with profoundly impaired immunoglobulin class-switch recombi-
nation. Nat. Immunol. 4, 1023 – 1028.
Intano, G.W., McMahan, C.A., Walter, R.B., McCarrey, J.R., Walter, C.A.,
2001. Mixed spermatogenic germ cell nuclear extracts exhibit high base
excision repair activity. Nucleic Acids Res. 29, 1366 – 1372.
Irvine, D.S., Twigg, J.P., Gordon, E.L., Fulton, N., Milne, P.A., Aitken,
R.J., 2000. DNA integrity in human spermatozoa: relationships with
semen quality. J. Androl. 21, 33 – 44.
Jansen, J., Olsen, A.K., Wiger, R., Naegeli, H., de Boer, P., van Der,
H.F., Holme, J.A., Brunborg, G., Mullenders, L., 2001. Nucleotide
excision repair in rat male germ cells: low level of repair in intact
cells contrasts with high dual incision activity in vitro. Nucleic Acids
Res. 29, 1791 – 1800.
Ji, B.T., Shu, X.O., Linet, M.S., Zheng, W., Wacholder, S., Gao, Y.T., Ying,
D.M., Jin, F., 1997. Paternal cigarette smoking and the risk of childhood
cancer among offspring of nonsmoking mothers. J. Natl. Cancer Inst.
89, 238 – 244.
Kim, N.K., Lee, S.H., Sohn, T.J., Roy, R., Mitra, S., Chung, H.M., Ko, J.J.,
Cha, K.Y., 2000. Spatial expression of a DNA repair gene, N-
methylpurine-DNA glycosylase (MPG) during development in mice.
Anticancer Res. 20, 3037 – 3043.
Klungland, A., Rosewell, I., Hollenbach, S., Larsen, E., Daly, G., Epe, B.,
Seeberg, E., Lindahl, T., Barnes, D.E., 1999. Accumulation of
premutagenic DNA lesions in mice defective in removal of oxidative
base damage. Proc. Natl. Acad. Sci. U.S.A. 96, 13300 – 13305.
Leadon, S.A., Cooper, P.K., 1993. Preferential repair of ionizing radiation-
induced damage in the transcribed strand of an active human gene is
defective in Cockayne syndrome. Proc. Natl. Acad. Sci. U.S.A. 90,
10499 – 10503.
Le Page, F., Kwoh, E.E., Avrutskaya, A., Gentil, A., Leadon, S.A., Sarasin,
A., Cooper, P.K., 2000. Transcription-coupled repair of 8-oxoguanine:
requirement for XPG, TFIIH, and CSB and implications for Cockayne
syndrome. Cell 101, 159 – 171.
Li, L., Peterson, C., Legerski, R., 1996. Sequence of the mouse XPC cDNA
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521– S531S530
and genomic structure of the human XPC gene. Nucleic Acids Res. 24,
1026 – 1028.
Lindahl, T., 1974. An N-glycosidase from Escherichia coli that releases
free uracil from DNA containing deaminated cytosine residues. Proc.
Natl. Acad. Sci. U.S.A. 71, 3649– 3653.
Missura, M., Buterin, T., Hindges, R., Hubscher, U., Kasparkova, J.,
Brabec, V., Naegeli, H., 2001. Double-check probing of DNA bending
and unwinding by XPA-RPA: an architectural function in DNA repair.
EMBO J. 20, 3554 – 3564.
Mitchell, J.R., Hoeijmakers, J.H., Niedernhofer, L.J., 2003. Divide and
conquer: nucleotide excision repair battles cancer and ageing. Curr.
Opin. Cell Biol. 15, 232 – 240.
Morris, I.D., Ilott, S., Dixon, L., Brison, D.R., 2002. The spectrum of DNA
damage in human sperm assessed by single cell gel electrophoresis
(Comet assay) and its relationship to fertilization and embryo develop-
ment. Hum. Reprod. 17, 990 – 998.
Nairn, R.S., Mitchell, D.L., Adair, G.M., Thompson, L.H., Siciliano, M.J.,
Humphrey, R.M., 1989. UV mutagenesis, cytotoxicity and split-dose
recovery in a human – CHO cell hybrid having intermediate (6-4)
photoproduct repair. Mutat. Res. 217, 193 – 201.
Nilsen, H., Rosewell, I., Robins, P., Skjelbred, C.F., Andersen, S.,
Slupphaug, G., Daly, G., Krokan, H.E., Lindahl, T., Barnes, D.E.,
2000. Uracil-DNA glycosylase (UNG)-deficient mice reveal a primary
role of the enzyme during DNA replication. Mol. Cell 5, 1059– 1065.
Nilsen, H., Stamp, G., Andersen, S., Hrivnak, G., Krokan, H.E.,
Lindahl, T., Barnes, D.E., 2003. Gene-targeted mice lacking the
Ung uracil-DNA glycosylase develop B-cell lymphomas. Oncogene
22, 5381 – 5386.
Nishioka, K., Ohtsubo, T., Oda, H., Fujiwara, T., Kang, D., Sugimachi, K.,
Nakabeppu, Y., 1999. Expression and differential intracellular local-
ization of two major forms of human 8-oxoguanine DNA glycosylase
encoded by alternatively spliced OGG1 mRNAs. Mol. Biol. Cell 10,
1637 – 1652.
Nouspikel, T., Hanawalt, P.C., 2002. DNA repair in terminally differ-
entiated cells. DNA Repair (Amst) 1, 59 75.
Oldereid, N.B., Angelis, P.D., Wiger, R., Clausen, O.P., 2001. Expression of
Bcl-2 family proteins and spontaneous apoptosis in normal human
testis. Mol. Hum. Reprod. 7, 403– 408.
Olsen, A.K., Bjortuft, H., Wiger, R., Holme, J., Seeberg, E., Bjoras, M.,
Brunborg, G., 2001. Highly efficient base excision repair (BER) in
human and rat male germ cells. Nucleic Acids Res. 29, 1781 – 1790.
Olsen, A.K., Duale, N., Bjoras, M., Larsen, C.T., Wiger, R., Holme, J.A.,
Seeberg, E.C., Brunborg, G., 2003. Limited repair of 8-hydroxy-7,8-
dihydroguanine residues in human testicular cells. Nucleic Acids Res.
31, 1351 – 1363.
Potts, R.J., Newbury, C.J., Smith, G., Notarianni, L.J., Jefferies, T.M., 1999.
Sperm chromatin damage associated with male smoking. Mutat. Res.
423, 103 – 111.
Prasad, R., Singhal, R.K., Srivastava, D.K., Molina, J.T., Tomkinson, A.E.,
Wilson, S.H., 1996. Specific interaction of DNA polymerase beta and
DNA ligase I in a multiprotein base excision repair complex from
bovine testis. J. Biol. Chem. 271, 16000 – 16007.
Radicella, J.P., Dherin, C., Desmaze, C., Fox, M.S., Boiteux, S., 1997.
Cloning and characterization of hOGG1, a human homolog of the
OGG1 gene of Saccharomyces cerevisiae. Proc. Natl. Acad. Sci.
U.S.A. 94, 8010 – 8015.
Rosenquist, T.A., Zharkov, D.O., Grollman, A.P., 1997. Cloning and
characterization of a mammalian 8-oxoguanine DNA glycosylase. Proc.
Natl. Acad. Sci. U.S.A. 94, 7429– 7434.
Shannon, M., Lamerdin, J.E., Richardson, L., McCutchen-Maloney, S.L.,
Hwang, M.H., Handel, M.A., Stubbs, L., Thelen, M.P., 1999. Character-
ization of the mouse Xpf DNA repair gene and differential expression
during spermatogenesis. Genomics 62, 427 – 435.
Sharma, R.K., Said, T., Agarwal, A., 2004. Sperm DNA damage and its
clinical relevance in assessing reproductive outcome. Asian J. Androl.
6, 139 – 148.
Shen, H.M., Chia, S.E., Ni, Z.Y., New, A.L., Lee, B.L., Ong, C.N., 1997.
Detection of oxidative DNA damage in human sperm and the
association with cigarette smoking. Reprod. Toxicol. 11, 675 – 680.
Sotomayor, R.E., Sega, G.A., 2000. Unscheduled DNA synthesis assay in
mammalian spermatogenic cells: an update. Environ. Mol. Mutagen. 36,
255 – 265.
Spierings, D.C., de Vries, E.G., Vellenga, E., de Jong, S., 2003. The
attractive Achilles heel of germ cell tumours: an inherent sensitivity to
apoptosis-inducing stimuli. J. Pathol. 200, 137– 148.
Sun, J.G., Jurisicova, A., Casper, R.F., 1997. Detection of deoxyribonucleic
acid fragmentation in human sperm: correlation with fertilization in
vitro. Biol. Reprod. 56, 602 – 607.
Thompson, M.J., Abdul-Rahman, S., Baker, T.G., Rafferty, J.A., Margison,
G.P., Bibby, M.C., 2000. Role of O6-alkylguanine-DNA alkyltransfer-
ase in the resistance of mouse spermatogenic cells to O6-alkylating
agents. J. Reprod. Fertil. 119, 339 – 346.
Tornaletti, S., Hanawalt, P.C., 1999. Effect of DNA lesions on transcription
elongation. Biochimie 81, 139 – 146.
Twigg, J.P., Irvine, D.S., Aitken, R.J., 1998. Oxidative damage to
DNA in human spermatozoa does not preclude pronucleus
formation at intracytoplasmic sperm injection. Hum. Reprod. 13,
1864 – 1871.
van der Spek, P.J., Visser, C.E., Hanaoka, F., Smit, B., Hagemeijer, A.,
Bootsma, D., Hoeijmakers, J.H., 1996. Cloning, comparative map-
ping, and RNA expression of the mouse homologues of the
Saccharomyces cerevisiae nucleotide excision repair gene RAD23.
Genomics 31, 20 – 27.
van Loon, A.A., Den Boer, P.J., van der Schans, G.P., Mackenbach,
P., Grootegoed, J.A., Baan, R.A., Lohman, P.H., 1991. Immuno-
chemical detection of DNA damage induction and repair at
different cellular stages of spermatogenesis of the hamster after in
vitro or in vivo exposure to ionizing radiation. Exp. Cell Res. 193,
303 – 309.
van Loon, A.A., Sonneveld, E., Hoogerbrugge, J., van der Schans, G.P.,
Grootegoed, J.A., Lohman, P.H., Baan, R.A., 1993. Induction and repair
of DNA single-strand breaks and DNA base damage at different cellular
stages of spermatogenesis of the hamster upon in vitro exposure to
ionizing radiation. Mutat. Res. 294, 139– 148.
Wood, R.D., 1997. Nucleotide excision repair in mammalian cells. J. Biol.
Chem. 272, 23465 – 23468.
Zenzes, M.T., Bielecki, R., Reed, T.E., 1999a. Detection of benzo(a)pyrene
diol epoxide-DNA adducts in sperm of men exposed to cigarette smoke.
Fertil. Steril. 72, 330 – 335.
Zenzes, M.T., Puy, L.A., Bielecki, R., Reed, T.E., 1999b. Detection of
benzo[a]pyrene diol epoxide-DNA adducts in embryos from smoking
couples: evidence for transmission by spermatozoa. Mol. Hum. Reprod.
5, 125 – 131.
Zitzmann, M., Rolf, C., Nordhoff, V., Schrader, G., Rickert-Fohring, M.,
Gassner, P., Behre, H.M., Greb, R.R., Kiesel, L., Nieschlag, E., 2003.
Male smokers have a decreased success rate for in vitro fertilization
and intracytoplasmic sperm injection. Fertil. Steril. 79 (Suppl. 3),
1550 – 1554.
A.-K. Olsen et al. / Toxicology and Applied Pharmacology 207 (2005) S521 S531 S531
... Males in spermiation have a high proportion of mature spermatozoa, which exclusively rely on antioxidants in seminal plasma to neutralize oxidative compounds (Figueroa et al., 2020). Oxidative compounds, such as reactive oxygen species (ROS), damage sperm DNA (Aitken et al., 2013), which cannot be repaired due to the extrusion of cytoplasmic organelles (Olsen et al., 2005). Moreover, Billard and Takashima (1983) reported that spermatozoa stored for an extended period in the sperm duct started degenerating by breaking down chromatin into filaments. ...
... Apoptosis plays a crucial role in spermatogenetic quality control [1]. DNA repair capacity of mammalian sperm declines in late spermatogenesis [2], a fraction of sperm accumulates DNA fragmentation. While at present, intracytoplasmic sperm injection (ICSI) is a major fertilization procedure in assisted reproductive technology (ART), many cohort studies have reported DNA fragmentation in human sperm nuclei as a major risk factor for sperm-derived congenital anomalies in ART [3][4][5]. ...
Preprint
Full-text available
We here developed a novel angle-modulated two-dimensional single cell pulsed-field gel electrophoresis (2D-SCPFGE). Variations in current-application-time and rotation angle generated different alignments of DNA fibers and segments. After the first run, the specimen was turned by 150 degrees (2D-SCPFGE-0-150) to detect naturally occurring the earliest stage of DNA fragmentation or 75 degrees (2D-SCPFGE-0-75) to analyze artificially induced cleavage. The former revealed that a part of long chain fibers remained at the origin and long segments were still tangled in the bundle of elongated fibers after the first run. The latter visualized the dose-dependent cleavage of DNA by EcoR1. Multicycle 2D-SCPFGE was useful for generating 2D-alignments of single nuclear DNA fibers, which is the first step for visualization of single-strand breaks on stretched fibers. To date, many articles have accepted the pathogenetic significances of DNA fragmentation in human sperm for male infertility and congenital anomaly. It is necessary to perform multivariate analyses of not only earliest-stage DNA fragmentation but also other types of damage, including single-strand breaks, in sequential DNA fibers. 2D-SCPFGE is the fundamental tool for understanding single nuclear DNA damages.
... It is generally acknowledged that DNA breaks are permissively generated and accumulated in mature sperm due to exposure to environmental mutagens, the haploid nature of the genome, an inaccessibility to DNA repair proteins, or dynamic changes in the DNA topology (Olsen et al. 2005;González-Marín et al. 2012). Our finding herein of a correlation between breakpoints and the peak of the ATAC-seq data, which indicates accessible chromatin regions, might be supported by the breakpoints distribution of mouse evolutionary genomic rearrangements to the accessible region in post-meiotic cells (Álvarez-González et al. 2022). ...
Article
Full-text available
Constitutional complex chromosomal rearrangements (CCRs) are rare cytogenetic aberrations arising in the germline via an unknown mechanism. Here we analyzed the breakpoint junctions of microscopically three-way or more complex translocations using comprehensive genomic and epigenomic analyses. All of these translocation junctions showed submicroscopic genomic complexity reminiscent of chromothripsis. The breakpoints were clustered within small genomic domains with junctions showing microhomology or microinsertions. Notably, all of the de novo cases were of paternal origin. The breakpoint distributions corresponded specifically to the ATAC-seq (assay for transposase-accessible chromatin with sequencing) read data peak of mature sperm and not to other chromatin markers or tissues. We propose that DNA breaks in CCRs may develop in an accessible region of densely packaged chromatin during post-meiotic spermiogenesis. Supplementary Information The online version contains supplementary material available at 10.1007/s00439-023-02591-9.
... The sperm formation process includes three phases: spermatogonia, meiosis, and spermiogenesis. The duration of the three phases is approximately the same (154). According to related studies, there are certain difficulties in achieving natural conception when the DNA damage rate of sperm is between 30% and 50%. ...
Article
Full-text available
In the complex and dynamic processes of replication, transcription, and translation of DNA molecules, a large number of replication errors or damage can occur which lead to obstacles in the development process of germ cells and result in a decreased reproductive rate. DNA damage repair has attracted widespread attention due to its important role in the maintenance and regulation of germ cells. This study reports on a systematic review of the role and mechanism of DNA damage repair in germline development. First, the causes, detection methods, and repair methods of DNA damage, and the mechanism of DNA damage repair are summarized. Second, a summary of the causes of abnormal DNA damage repair in germ cells is introduced along with common examples, and the relevant effects of germ cell damage. Third, we introduce the application of drugs related to DNA damage repair in the treatment of reproductive diseases and related surgical treatment of abnormal DNA damage, and summarize various applications of DNA damage repair in germ cells. Finally, a summary and discussion is given of the current deficiencies in DNA damage repair during germ cell development and future research development. The purpose of this paper is to provide researchers engaged in relevant fields with a further systematic understanding of the relevant applications of DNA damage repair in germ cells and to gain inspiration from it to provide new research ideas for related fields.
... Letters are presenting statistically significant (p < 0.05) differences between subcapsular (SC) areas of control and VCL groups; * and ** are presenting the significant differences (p < 0.05) between central zones of control and VCL group, and # is presenting significant differences (p < 0.05) between subcapsular and central zones in each group testicular cellularity. Concerning the BER pathway's role in repairing the oxidative DNA damage as well as minimizing the number of potential mutations on DNA after oxidative damage [54][55][56], we can suggest that VCL by suppressing the UNG and MPG glycosylases expression can boost the nucleophilic DNA bases oxidation leading to base loss and/ or strand breaks. In line with this suggestion, as mentioned above, we found an increased 8-oxo-dG reactivity representing guanine oxidation as well as elevated late apoptotic cells, exhibiting VCL-induced DNA strand break. ...
Article
Varicocele (VCL) has been shown to induce severe oxidative stress in the testicular tissue resulting in 35% of males with primary infertility. To compare the exacerbating impacts of varicose on oxidative DNA damage and homeostatic antioxidant reactions in the seminiferous tubules (ST), enclosed and far from varicose vessels. Thirty mature Wistar rats were divided into control and VCL-induced groups. To approve VCL, the testicular diameters, volume, and blood circulation were measured using B-mode and Doppler ultrasonography. Next, to confirm oxidative stress (OS), the global homeostatic antioxidant biomarkers were evaluated. Moreover, the OS-induced oxidative DNA damage and homeostatic antioxidant reactions were compared between STs nearby and far from varicose vessels. Finally, to clarify the DNA damage-induced impact on the cell cycle progression, the global and local expressions of Cyclin D1, Cdk4, and p21 were examined. The VCL-induced group exhibited diminished global antioxidant status (marked with TAC, GPX, SOD, and CAT) and UNG and MPG expression levels. Moreover, the cross-sections of the VCL group represented a prominent reduction in the UNG, MPG, Cyclin D1, and cdk4, and upregulation in the p21 expression levels, more prominently in the STs nearby varicose vessels. Concerning severe oxidative DNA damage and intensive molecular changes in the STs nearby the varicose vessels, they can be considered the main cause of oxidative DNA damage in enclosed tubules. Thus, the varicose-mediated oxidative DNA damage negatively impacts the cell cycle progression in the tubules more intensively in the subcapsular area.
... The OECD suggests that protocol modifications and validation studies are necessary before TG 489 can be revised to include an analysis of testicular germ cells. Nevertheless, the comet assay has been used to measure DNA damage in gonadal cells in both research settings (Asare et al., 2016;Bjorge et al., 1995;1996a;Brunborg et al., 2015;Olsen et al., 2001;Collins, 2004;Olsen et al., 2005;Olive & Banath, 2006;Speit & Hartmann, 2006;Dhawan et al., 2009;Hansen et al., 2010;Graupner et al., 2014;2015;Gutzkow et al., 2016Graupner et al., 2017Sharma et al., 2018) and in a regulatory perspective (Brendler-Schwaab et al., 2005;Vasquez, 2012;Graupner et al., 2014;Frotschl, 2015;Koppen et al., 2017). ...
Article
Full-text available
The in vivo comet assay is widely used to measure genotoxicity; however, the current OECD test guideline (TG 489) does not recommend using the assay to assess testicular germ cells, due to the presence of testicular somatic cells. An adapted approach to specifically assess testicular germ cells within the comet assay is certainly warranted, considering regulatory needs for germ-cell specific genotoxicity data in relation to the increasing global production of and exposure to potentially hazardous chemicals. Here we provide a proof-of-concept to selectively analyze round spermatids and primary spermatocytes, distinguishing them from other cells of the testicle. Utilizing the comet assay recordings of DNA content (total fluorescence intensity) and DNA damage (% tail intensity) of individual comets, we developed a framework to distinguish testicular cell populations based on differences in DNA content/ploidy and appearance. Haploid round spermatid comets are identified through 1) visual inspection of DNA content distributions, 2) setting DNA content thresholds, and 3) modelling DNA content distributions using a normal mixture distribution function. We also describe an approach to distinguish primary spermatocytes during comet scoring, based on their high DNA content and large physical size. Our concept allows both somatic and germ cells to be analyzed in the same animal, adding a versatile, sensitive, rapid, and resource efficient assay to the limited genotoxicity assessment toolbox for germ cells. An adaptation of TG 489 facilitates accumulation of valuable information regarding distribution of substances to germ cells and their potential for inducing germ cell gene mutations and structural chromosomal aberrations. This article is protected by copyright. All rights reserved.
... 11 In fact, the ability to repair DNA lesions declines dramatically during final stages of spermatogenesis. 12 Meiotic interstrand DNA damage that escapes paternal repair can cause chromosomal aberrations in the zygote by maternal misrepair taking place after fertilisation. 13 In addition to the physiological, recombinationrelated DNA breakage, DNA of the developing male germ cell can undergo SSB or DSB through different pathological mechanisms, especially those related to excessive production or insufficient scavenging of reactive oxygen species (ROS). ...
Article
Full-text available
DNA of human spermatozoa can be subject to various kinds of modifications acquired throughout life. Put simply, two basic types of acquired sperm DNA modifications can be distinguished: genetic and epigenetic. Genetic modifications cause alterations of the DNA sequence and mainly result from the formation of breakpoints leading to sperm DNA fragmentation. Epigenetic modifications include a vast spectrum of events that influence the expression of different genes without altering their DNA sequence. Both the genetic and the epigenetic modifications of sperm DNA can negatively influence embryonic development, cause miscarriages, and be the origin of different health problems for the offspring. As to sperm DNA fragmentation, reliable diagnostic methods are currently available. On the other hand, the detection of potentially harmful epigenetic modifications in spermatozoa is a much more complicated issue. Different treatment options can be chosen to solve problems associated with sperm DNA fragmentation. Some are relatively simple and noninvasive, based on oral treatments with antioxidants and other agents, depending on the underlying cause. In other cases, the recourse to different micromanipulation-assisted in vitro fertilisation techniques is necessary to select spermatozoa with minimal DNA damage to be injected into oocytes. The treatment of cases with epigenetic DNA modifications is still under investigation. Preliminary data suggest that some of the techniques used in cases of extensive DNA fragmentation can also be of help in those of epigenetic modifications; however, further progress will depend on the availability of more reliable diagnostic methods with which it will be possible to evaluate the effects of different therapeutic interventions.
Article
Full-text available
We here developed a novel angle-modulated two-dimensional single cell pulsed-field gel electrophoresis (2D-SCPFGE). Variations in current-application-time and rotation angle generated different alignments of DNA fibers and segments. After the first run, the specimen was turned by 150° (2D-SCPFGE-0–150) to detect naturally occurring the earliest stage of DNA fragmentation or 75° (2D-SCPFGE-0–75) to analyze artificially induced cleavage. The former revealed that a part of long chain fibers remained at the origin and long segments were still tangled in the bundle of elongated fibers after the first run. The latter visualized the dose-dependent cleavage of DNA by EcoR1. Multicycle 2D-SCPFGE was useful for generating 2D-alignments of single nuclear DNA fibers, which is the first step for visualization of single-strand breaks on stretched fibers. To date, many articles have accepted the pathogenetic significances of DNA fragmentation in human sperm for male infertility and congenital anomaly. It is necessary to perform multivariate analyses of not only earliest-stage DNA fragmentation but also other types of damage, including single-strand breaks, in sequential DNA fibers. 2D-SCPFGE is the fundamental tool for understanding single nuclear DNA damages.
Poster
Full-text available
Introduction Xeroderma pigmentosum (XP) is a rare, autosomal recessive disorder that results in deficient nucleotide excision repair (NER) of DNA repair of ultraviolet (UV)-induced lesions. Patients with XP are at increased susceptibility to skin malignancies and other dermal complications. Despite no known cure, various therapeutic modalities have been investigated to help improve quality of life and prevent the progression of XP. However, with the recent advances in the field of XP therapies, a comprehensive and critical review of the available literature regarding the effectiveness and future potential of these interventions is warranted. This systematic review aims to address and evaluate the current state of XP therapeutics, encompassing photoprotection, topical and systemic agents, and surgical interventions. Our analysis highlights promising treatments to manage XP and discusses opportunities in the field for further research. We highlight the challenges in clinical trial design for this rare and complex disease, which include a small patient population, ethical considerations, and clinical heterogeneity. Methods A systematic review of the literature was conducted through EMBASE/Elsevier, Scopus, Medline, PubMed, BVS, SciELO, and Lilacs, following the preferred reporting items for systematic reviews and meta-analyses (PRISMA) guidelines. The search strategy was as follows: (("xeroderma pigmentosum") OR (("xeroderma") AND ("pigmentosum"))) AND (("therapy") OR ("therapeutic") OR ("treatment") OR ("medication") OR ("intervention") OR ("drug") OR ("surgery") OR ("surgical") OR ("operative") OR ("procedure")). Additionally, a manual search was conducted through the connected authors reference lists for relevant sources. Results Our analysis revealed that photoprotection measures remain the cornerstone for XP management, including sun avoidance and protective clothing. The use of DNA repair enzymes (e.g., T4 endonuclease V) and topical agents (e.g., imiquimod and 5-fluorouracil) helped to significantly reduce new actinic keratoses and basal cell carcinomas. Vitamin supplementation in the form of Nicotinamide showed promising results over a 12-month period reducing the incidence of basal cell carcinoma by 20% and squamous cell carcinoma by 30%. Surgical interventions and cryotherapy of precancerous lesions or skin grafting, demonstrating promising results in select cases. Conclusions XP is a complex disorder which requires a multifaceted approach emphasising the importance of early intervention through photoprotection and individualised therapeutic regimens. The lack of available information on longitudinal impacts of surgical and therapeutic interventions calls for further research in this field.
Book
Voortplanting en genetica zijn de centrale thema’s van de biologie. Ook hebben ze te maken met levensvragen als: waar komen we vandaan, wat zijn we en waar gaan we naartoe. In de gecombineerde wetenschap, voorplantingsgenetica, worden de generaties met elkaar verbonden en speelt vruchtbaarheid een grote rol. De cellijn die daar verantwoordelijk voor is, staat bekend als de kiembaan, met de gameten als de functionele producten. Het boek volgt de kiembaan en poogt door een vrijer en meer verhalend gebruik van de taal de kloof tussen een academisch tekstboek en de maatschappij te overbruggen. Persoonlijke ervaringen worden niet gemeden, maar domineren niet. Het boek is bedoeld voor mensen die professioneel met deze materie te maken hebben of krijgen, ook in het onderwijs en de journalistiek, maar is door de prachtige kleurenillustraties en een actuele verklarende woordenlijst ook geschikt voor geïnteresseerden zonder professionele achtergrond.
Article
Full-text available
TheSaccharomyces cerevisiae RAD23gene is involved in nucleotide excision repair (NER). Two human homologs ofRAD23, HHR23AandHHR23B(HGMW-approved symbols RAD23A and RAD23B), were previously isolated. The HHR23B protein is complexed with the protein defective in the cancer-prone repair syndrome xeroderma pigmentosum, complementation group C, and is specifically involved in the global genome NER subpathway. The cloning of both mouse homologs (designated MHR23A and MHR23B) and detailed sequence comparison permitted the deduction of the following overall structure for all RAD23 homologs: an ubiquitin-like N-terminus followed by a strongly conserved 50-amino-acid domain that is repeated at the C-terminus. We also found this domain as a specific C-terminal extension of one of the ubiquitin-conjugating enzymes, providing a second link with the ubiquitin pathway. By means ofin situhybridization,MHR23Awas assigned to mouse chromosome 8C3 andMHR23Bto 4B3. Because of the close chromosomal proximity of humanXPCandHHR23B,the mouseXPCchromosomal location was determined (6D). Physical disconnection of the genes in mouse argues against a functional significance of the colocalization of these genes in human. Northern blot analysis revealed constitutive expression of bothMHR23genes in all tissues examined. Elevated RNA expression of bothMHR23genes was observed in testis. Although the RAD23 equivalents are well conserved during evolution, the mammalian genes did not express the UV-inducible phenotype of their yeast counterpart. This may point to a fundamental difference between the UV responses of yeast and human. No stage-specific mRNA expression during the cell cycle was observed for the mammalianRAD23homologs.
Article
Full-text available
The multiprotein factor composed of XPA and replication protein A (RPA) is an essential subunit of the mammalian nucleotide excision repair system. Although XPA–RPA has been implicated in damage recognition, its activity in the DNA repair pathway remains controversial. By replacing DNA adducts with mispaired bases or non-hybridizing analogues, we found that the weak preference of XPA and RPA for damaged substrates is entirely mediated by indirect readout of DNA helix conformations. Further screening with artificially distorted substrates revealed that XPA binds most efficiently to rigidly bent duplexes but not to single-stranded DNA. Conversely, RPA recognizes single-stranded sites but not backbone bending. Thus, the association of XPA with RPA generates a double-check sensor that detects, simultaneously, backbone and base pair distortion of DNA. The affinity of XPA for sharply bent duplexes, characteristic of architectural proteins, is not compatible with a direct function during recognition of nucleotide lesions. Instead, XPA in conjunction with RPA may constitute a regulatory factor that monitors DNA bending and unwinding to verify the damage-specific localization of repair complexes or control their correct three-dimensional assembly.
Article
The objective of this study was to determine the incidence of DNA fragmentation in human sperm, and to correlate any detected DNA damage with semen analysis parameters and fertilization rates in in vitro fertilization (IVF). A total of 298 semen samples were collected from men in the infertility program at The Toronto Hospital. For each sample, the percentage of sperm with DNA fragmentation was determined using the method of terminal deoxynucleotidyl transferase-mediated dUTP-biotin end-labeling (TUNEL) and fluorescence-activated cell sorting. The percentage of sperm with fragmented DNA was less than 4% in the majority of samples but ranged from 5% to 40% in approximately 27% of the samples. A negative correlation was found between the percentage of DNA fragmentation and the motility, morphology, and concentration of the ejaculated sperm. In 143 IVF samples, a significant negative association was also found between the percentage of sperm with DNA fragmentation and fertilization rate (p = 0.008) and embryo cleavage rate (p = 0.91). In addition, 35 men who smoked demonstrated an increased percentage of sperm with fragmented DNA (4.7 ± 1.2%) as compared to 78 nonsmokers (1.1 ± 0.2%; p = 0.01). These results demonstrate a negative association between semen analysis parameters and sperm with fragmented DNA. Since extremely poor semen samples are the indication for intracytoplasmic sperm injection, there is a high likelihood that sperm with fragmented DNA may be selected by chance and used for oocyte injection, resulting in poor fertilization and/or cleavage rates.
Article
In normal human cells, damage due to ultraviolet light is preferentially removed from active genes by nucleotide excision repair (NER) in a transcription-coupled repair (TCR) process that requires the gene products defective in Cockayne syndrome (CS). Oxidative damage, including thymine glycols, is shown to be removed by TCR in cells from normal individuals and from xeroderma pigmentosum (XP)-A, XP-F, and XP-G patients who have NER defects but not from XP-G patients who have severe CS. Thus, TCR of oxidative damage requires an XPG function distinct from its NER endonuclease activity. These results raise the possibility that defective TCR of oxidative damage contributes to the developmental defects associated with CS.
Article
Tobacco smoking is deleterious to reproduction. Benzo[a]pyrene (B[a]P) is a potent carcinogen in cigarette smoke. Its reactive metabolite induces DNA-adducts, which can cause mutations. We investigated whether B[a]P diol epoxide (BPDE) DNA adducts are detectable in preimplantation embryos in relation to parental smoking. A total of 17 couples were classified by their smoking habits: (i) both partners smoke; (ii) wife nonsmoker, husband smokes; and (iii) both partners were non-smokers. Their 27 embryos were exposed to an anti-BPDE monoclonal antibody that recognizes BPDE‐DNA adducts. Immunostaining was assessed in each embryo and an intensity score was calculated for embryos in each smoking group. The proportion of blastomeres which stained was higher for embryos of smokers than for non-smokers (0.723 versus 0.310). The mean intensity score was also higher for embryos of smokers (1.40 K 0.28) than for non-smokers (0.38 K 0.14; P J 0.015), but was similar for both types of smoking couples. The mean intensity score was positively correlated with the number of cigarettes smoked by fathers (P J 0.02). Increased mean immunostaining in embryos from smokers, relative to non-smokers, indicates a relationship with parental smoking. The similar levels of immunostaining in embryos from both types of smoking couples suggest that transmission of modified DNA is mainly through spermatozoa. We confirmed paternal transmission of modified DNA by detection of DNA adducts in spermatozoa of a smoker father and his embryo.
Article
This paper has been retracted by the authors.Reason: In a letter to the Editor dated May 11, 2005, S.A.L. requested that Figures 1 and 2 containing his data be retracted, since examination of the raw data revealed that “there was an error in experiments that could have influenced, if not accounted for, the results presented”. Examination by P.K.C. of the original data from the Leadon laboratory for these two figures revealed that unequal amounts of DNA had been loaded for analysis of repair in the transcribed and non-transcribed strands, with the unequal loading being sufficient to account for the appearance of preferential repair of the transcribed strand. These details in the conduct of the assay occurred without the knowledge of the other authors. Since key aspects of the data in this paper cannot be reproduced, the experimental evidence presented no longer provides sufficient support for the main conclusions. Thus, the authors are retracting the paper and apologize for any inconvenience this may have caused the scientific community.
Article
Testicular germ cell tumours (TGCTs) are extremely sensitive to cisplatin-containing chemotherapy. The rapid time course of apoptosis induction after exposure to cisplatin suggests that TGCT cells are primed to undergo programmed cell death as an inherent property of the cell of origin. In fact, apoptosis induction of germ cells in the testis is an important physiological mechanism to control the quality and quantity of the gametes produced. Although p53 protein is highly expressed in the majority of TGCTs, almost no p53 mutations have been detected. Interestingly, p53 overexpression is associated with loss of p21 and gain of mdm2 expression, which might indicate a partial loss in functionality of the p53 regulatory pathway in TGCTs. Besides p21, TGCTs often show low expression of other proteins involved in the regulation of cell cycle progression, such as the retinoblastoma protein and members of the INK4 family. It can be postulated that the deregulated G1-S phase checkpoint results in premature entry into the S phase upon DNA damage. In addition to Bcl-2 family members that are involved in the regulation of germ cell apoptosis in the normal testis via the mitochondrial death pathway, the Fas death pathway is also known to regulate apoptosis of germ cells in the testis. Since chemotherapy has been shown to activate the Fas death pathway and TGCTs co-express both Fas and its ligand FasL, TGCT cells might undergo apoptosis upon cisplatin treatment via autocrine or paracrine activation of the Fas system by FasL. The hypothesis suggested here is that the lack of cell cycle arrest following a cisplatin-containing treatment, together with the activation of the Fas death pathway and the mitochondrial death pathway, explains the rapid and efficient apoptosis of TGCT cells. Defining the mechanisms involved in the cisplatin sensitivity of TGCTs will provide tools to increase cisplatin sensitivity in other human tumours with acquired or intrinsic resistance.
Article
An immunochemical method has been used to detect quantitatively DNA damage caused by ionizing radiation in germ cells. With this method, DNA strand breaks as well as lesions converted into breaks in alkaline medium are measured as a function of controlled partial unwinding of the DNA, a time-dependent process starting at each breakage site, followed by the determination of the relative amount of single-stranded regions by use of a single-strand specific monoclonal antibody. With this method the induction and repair of DNA damage in different cellular stages of spermatogenesis (spermatocytes, round and elongated spermatids) of the hamster were investigated. Germ cells were irradiated in vitro with 60Co-gamma-rays, at doses between 0 and 5 Gy. A linear dose-response relationship was observed. Spermatocytes and round spermatids had normal, fast repair of the lesions when compared with the repair of these sites in cultured V79 or CHO cells and human lymphocytes. The elongated spermatids, however, showed hardly any repair. Similar results were obtained after the in vivo gamma-irradiation of hamsters with doses of 0. 4, and 8 Gy and subsequent isolation of germ cells. The damage was still detectable in the elongated spermatids at 24 h after exposure. The results of the experiments show substantial differences in repair capacity between different stages of germ cell development. Because DNA is the major target for mutation induction, this assay may be useful for assessment of the genetic risk of exposure of male germ cells to ionizing radiation, in relation to the stage of development.
Article
Two types of DNA ligase, I and II, have been purified approximately 4,000-fold from mouse testes and 500-fold from nuclei of mouse spermatocytes. DNA ligase I and II consisted of single polypeptides with molecular weights of 95,000 and 65,000, respectively, according to the estimation by SDS-polyacrylamide gel electrophoresis and the AMP-binding assay. Ligase activities were higher in premeiotic spermatogonia and spermatocytes than those in liver and bone marrow cells. Moreover, DNA ligase II showed rapid increase during meiotic prophase and a decrease in round spermatids. Since this behavior of DNA ligase II is consistent with that of m-rec and DNA polymerase beta, both of which have been shown to be involved in DNA recombination in meiotic cells, DNA ligase II might be an enzyme which works at the final step of meiotic recombination reaction.
Article
Somatic cell hybrids constructed between UV-hypersensitive Chinese hamster ovary cell line UV20 and human lymphocytes were used to examine the influence of a human DNA repair gene, ERCC1, on UV photoproduct repair, mutability at several drug-resistance loci, UV cytotoxicity and UV split-dose recovery. In hybrid cell line 20HL21-4, which contains human chromosome 19, UV-induced mutagenesis at the APRT, HPRT and Na+/K+-ATPase loci was comparable to that in repair-proficient CHO AA8 cells, whereas cell line 20HL21-7, a reduced human-CHO hybrid not containing human chromosome 19, exhibited a hypermutable phenotype at all 3 loci indistinguishable from that of UV20 cells. The response of 20HL21-4 cells to UV cytotoxicity reflected substantial but incomplete restoration of wild-type UV cytotoxic response, whereas responses of UV20 and 20HL21-7 cell lines to UV cytotoxicity were essentially the same, reflecting several-fold UV hypersensitivity. Repair of UV-induced (5-6) cyclobutane dimers and (6-4) photoproducts was examined by radioimmunoassay; (6-4) photoproduct repair was deficient in UV20 and 20HL21-7 cell lines, and intermediate in 20HL21-4 cells relative to wild-type CHO AA8 cells. UV split-dose recovery in 20HL21-4 cells was also intermediate relative to AA8 cells. These results show that the human ERCC1 gene on chromosome 19 is responsible for substantial restoration of UV survival and mutation responses in repair-deficient UV20 cells, but only partially restores (6-4) UV photoproduct repair and UV split-dose recovery.