11
Mutagenic potential of nitrogen mustard-induced formamidopyrimidine DNA adduct: Contribution of the non-canonical -anomer Received for publication, June 15, 2017, and in revised form, September 15, 2017 Published, Papers in Press, September 28, 2017, DOI 10.1074/jbc.M117.802520 Irina G. Minko , Carmelo J. Rizzo § , and R. Stephen Lloyd ‡¶1 From the Oregon Institute of Occupational Health Sciences and the Departments of Molecular and Medical Genetics and Physiology and Pharmacology, Oregon Health & Science University, Portland, Oregon 97239 and the § Departments of Chemistry and Biochemistry, Vanderbilt-Ingram Cancer Center, Vanderbilt University, Nashville, Tennessee 37235 Edited by Patrick Sung Nitrogen mustards (NMs) are DNA-alkylating compounds that represent the earliest anticancer drugs. However, clinical use of NMs is limited because of their own mutagenic proper- ties. The mechanisms of NM-induced mutagenesis remain unclear. The major product of DNA alkylation by NMs is a cat- ionic NM-N7-dG adduct that can yield the imidazole ring- fragmented lesion, N 5 -NM-substituted formamidopyrimidine (NM-Fapy-dG). Characterization of this adduct is complicated because it adopts different conformations, including both a canonical - and an unnatural -anomeric configuration. Although formation of NM-Fapy-dG in cellular DNA has been demonstrated, its potential role in NM-induced mutagenesis is unknown. Here, we created site-specifically modified single- stranded vectors for replication in primate (COS7) or Escherichia coli cells. In COS7 cells, NM-Fapy-dG caused targeted mutations, predominantly G 3 T transversions, with overall frequencies of 11–12%. These frequencies were 2-fold higher than that induced by 8-oxo-dG adduct. Replication in E. coli was essentially error-free. To elucidate the mechanisms of bypass of NM-Fapy-dG, we performed replication assays in vitro with a high-fidelity DNA polymerase, Saccharomyces cerevisiae polymerase (pol) . It was found that pol could catalyze high-fidelity synthesis past NM- Fapy-dG, but only on a template subpopulation, presumably con- taining the -anomeric adduct. Consistent with the low mutagenic potential of the -anomer in vitro, the mutation frequency was significantly reduced when conditions for vector preparation were modified to favor this configuration. Collectively, these data impli- cate the -anomer as a major contributor to NM-Fapy-dG-induced mutagenesis in primate cells. Nitrogen mustards (NMs) 2 are bifunctional alkylating com- pounds derived from bis(2-chloroethyl)amine (Fig. 1A). Repre- senting the earliest of the anticancer drugs, these agents, par- ticularly cyclophosphamide and melphalan, are still employed for the treatment of hematologic malignancies and various solid tumors (1, 2) and also as components of conditioning reg- imens for hematopoietic stem cell transplantation (3). The therapeutic effect of NMs is attributed to their ability to form DNA interstrand cross-links and thus, interfere with critical biological processes that require separation of DNA strands (4). However, the clinical use of NMs is limited because they are themselves highly carcinogenic, with the incidence of second- ary malignancies, mostly acute myeloid leukemia, being signif- icant in patients treated with these agents (5, 6). Although the carcinogenic potential of NMs has been recognized for decades, DNA lesions that are primarily responsible for NM-induced mutations remain undetermined. Reaction of NMs with DNA yields a multitude of products. The most common site of alkylation is N7-dG (7–13) with the cationic monoadduct (N7-NM-dG) being the major initial spe- cies (Fig. 1B). Other known sites of modification include N1- and N3-dA (11, 12, 14, 15), N3-dC (15, 16), and O 6 -dG (17). Furthermore, the alkylated base can react with another DNA base (7, 10 –13, 15) or with a cysteine thiol within proteins (18, 19) to form DNA–DNA or DNA–protein cross-links, respec- tively. Despite the recognized clinical significance, DNA inter- strand cross-links represent only a small fraction of the NM-in- duced DNA lesions (13). In addition to cross-linking, the cationic N7-NM-dG adduct can undergo a secondary reaction that proceeds via the addition of a hydroxide to the C8 position with a subsequent imidazole ring-opening, yielding the N 5 -NM-substituted formamidopy- rimidine adduct (NM-Fapy-dG) (Fig. 1C). Although the pos- sibility of such ring fragmentation had been reported in the early 1980s (8 –10, 20), its biological relevance had remained uncertain until 2015 when Rizzo, Turesky, and colleagues (21) established an ion-trap multistage scan MS method to simulta- neously measure various dG adducts of bis(2-chloroethyl)eth- ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG adducts were formed at a level 20% that of the N7-NM-dG adducts. When these studies were extended to a tumor cell line that was exposed to bis(2-chloro- This work was supported by National Institutes of Health, NCI Grants T32 CA106195, P01 CA160032, and P30 CA068485 and NIEHS Grants P30 ES000267 and T32 ES0007060. The authors declare that they have no con- flicts of interest with the contents of this article. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. This article contains supplemental Figs. S1 and S2. 1 To whom correspondence should be addressed. Tel.: 503-494-9957; Fax: 503-494-6831; E-mail: [email protected]. 2 The abbreviations used are: NM, nitrogen mustard; Fapy, formamidopyrimi- dine; NM-Fapy-dG, N 5 -NM-substituted formamidopyrimidine dG; Me-Fapy- dG, N 5 -methyl substituted formamidopyrimidine dG; 8-oxo-dG, 8-oxo-7,8-di- hydro-2-dG; Endo IV, Endonuclease IV; pol, polymerase. cro ARTICLE 18790 J. Biol. Chem. (2017) 292(46) 18790 –18799 © 2017 by The American Society for Biochemistry and Molecular Biology, Inc. Published in the U.S.A. by guest on May 11, 2020 http://www.jbc.org/ Downloaded from

Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

Mutagenic potential of nitrogen mustard-inducedformamidopyrimidine DNA adduct: Contribution of thenon-canonical �-anomerReceived for publication, June 15, 2017, and in revised form, September 15, 2017 Published, Papers in Press, September 28, 2017, DOI 10.1074/jbc.M117.802520

Irina G. Minko‡, Carmelo J. Rizzo§, and R. Stephen Lloyd‡¶1

From the ‡Oregon Institute of Occupational Health Sciences and the ¶Departments of Molecular and Medical Genetics andPhysiology and Pharmacology, Oregon Health & Science University, Portland, Oregon 97239 and the §Departments of Chemistryand Biochemistry, Vanderbilt-Ingram Cancer Center, Vanderbilt University, Nashville, Tennessee 37235

Edited by Patrick Sung

Nitrogen mustards (NMs) are DNA-alkylating compoundsthat represent the earliest anticancer drugs. However, clinicaluse of NMs is limited because of their own mutagenic proper-ties. The mechanisms of NM-induced mutagenesis remainunclear. The major product of DNA alkylation by NMs is a cat-ionic NM-N7-dG adduct that can yield the imidazole ring-fragmented lesion, N5-NM-substituted formamidopyrimidine(NM-Fapy-dG). Characterization of this adduct is complicatedbecause it adopts different conformations, including both acanonical �- and an unnatural �-anomeric configuration.Although formation of NM-Fapy-dG in cellular DNA has beendemonstrated, its potential role in NM-induced mutagenesis isunknown. Here, we created site-specifically modified single-stranded vectors for replication in primate (COS7) or Escherichiacoli cells. In COS7 cells, NM-Fapy-dG caused targeted mutations,predominantly G 3 T transversions, with overall frequencies of�11–12%. These frequencies were �2-fold higher than thatinduced by 8-oxo-dG adduct. Replication in E. coli was essentiallyerror-free. To elucidate the mechanisms of bypass of NM-Fapy-dG,we performed replication assays in vitro with a high-fidelity DNApolymerase, Saccharomyces cerevisiae polymerase (pol) �. It wasfound that pol � could catalyze high-fidelity synthesis past NM-Fapy-dG, but only on a template subpopulation, presumably con-taining the �-anomeric adduct. Consistent with the low mutagenicpotential of the �-anomer in vitro, the mutation frequency wassignificantly reduced when conditions for vector preparation weremodified to favor this configuration. Collectively, these data impli-cate the �-anomer as a major contributor to NM-Fapy-dG-inducedmutagenesis in primate cells.

Nitrogen mustards (NMs)2 are bifunctional alkylating com-pounds derived from bis(2-chloroethyl)amine (Fig. 1A). Repre-

senting the earliest of the anticancer drugs, these agents, par-ticularly cyclophosphamide and melphalan, are still employedfor the treatment of hematologic malignancies and varioussolid tumors (1, 2) and also as components of conditioning reg-imens for hematopoietic stem cell transplantation (3). Thetherapeutic effect of NMs is attributed to their ability to formDNA interstrand cross-links and thus, interfere with criticalbiological processes that require separation of DNA strands (4).However, the clinical use of NMs is limited because they arethemselves highly carcinogenic, with the incidence of second-ary malignancies, mostly acute myeloid leukemia, being signif-icant in patients treated with these agents (5, 6). Although thecarcinogenic potential of NMs has been recognized for decades,DNA lesions that are primarily responsible for NM-inducedmutations remain undetermined.

Reaction of NMs with DNA yields a multitude of products.The most common site of alkylation is N7-dG (7–13) with thecationic monoadduct (N7-NM-dG) being the major initial spe-cies (Fig. 1B). Other known sites of modification include N1-and N3-dA (11, 12, 14, 15), N3-dC (15, 16), and O6-dG (17).Furthermore, the alkylated base can react with another DNAbase (7, 10 –13, 15) or with a cysteine thiol within proteins (18,19) to form DNA–DNA or DNA–protein cross-links, respec-tively. Despite the recognized clinical significance, DNA inter-strand cross-links represent only a small fraction of the NM-in-duced DNA lesions (13).

In addition to cross-linking, the cationic N7-NM-dG adductcan undergo a secondary reaction that proceeds via the additionof a hydroxide to the C8 position with a subsequent imidazolering-opening, yielding the N5-NM-substituted formamidopy-rimidine adduct (NM-Fapy-dG) (Fig. 1C). Although the pos-sibility of such ring fragmentation had been reported in theearly 1980s (8 –10, 20), its biological relevance had remaineduncertain until 2015 when Rizzo, Turesky, and colleagues (21)established an ion-trap multistage scan MS method to simulta-neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA invitro, the NM-Fapy-dG adducts were formed at a level �20%that of the N7-NM-dG adducts. When these studies wereextended to a tumor cell line that was exposed to bis(2-chloro-

This work was supported by National Institutes of Health, NCI Grants T32CA106195, P01 CA160032, and P30 CA068485 and NIEHS Grants P30ES000267 and T32 ES0007060. The authors declare that they have no con-flicts of interest with the contents of this article. The content is solely theresponsibility of the authors and does not necessarily represent the officialviews of the National Institutes of Health.

This article contains supplemental Figs. S1 and S2.1 To whom correspondence should be addressed. Tel.: 503-494-9957; Fax:

503-494-6831; E-mail: [email protected] The abbreviations used are: NM, nitrogen mustard; Fapy, formamidopyrimi-

dine; NM-Fapy-dG, N5-NM-substituted formamidopyrimidine dG; Me-Fapy-dG, N5-methyl substituted formamidopyrimidine dG; 8-oxo-dG, 8-oxo-7,8-di-hydro-2�-dG; Endo IV, Endonuclease IV; pol, polymerase.

croARTICLE

18790 J. Biol. Chem. (2017) 292(46) 18790 –18799

© 2017 by The American Society for Biochemistry and Molecular Biology, Inc. Published in the U.S.A.

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 2: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

ethyl)ethylamine, NM-Fapy-dG was detected at the level of 180adducts per 107 nucleosides, representing �13% of all adductsmeasured. DNA cross-links were also detected in the productsfrom both of the experimental systems described above. Inter-estingly, a fraction of these cross-links contained dGs thatunderwent ring-opening. These data demonstrated thatN7-NM-dG monoadducts or cross-links are routinely con-verted to the corresponding Fapy-dG species and thus, contri-bution of these ring-fragmented lesions to NM-inducedmutagenesis should be considered.

To facilitate investigations on the structure and function ofthe NM-Fapy-dG adducts, the Rizzo group developed chemis-try for the site-specific synthesis of oligodeoxynucleotides con-taining bis(2-chloroethyl)ethylamine-derived Fapy-dG (Fig.1A) and employed Escherichia coli endonuclease IV (Endo IV),an enzyme that selectively incises double-stranded DNA at�-nucleotides (22, 23) to probe the anomeric configuration ofNM-Fapy-dG (24). These data showed that similar to unsubsti-tuted (25) or methyl-substituted Fapy-dG (Me-Fapy-dG) (26),NM-Fapy-dG exists as a mixture of interconverting �- and�-anomeric forms (Fig. 1, C and D), with the former being pre-dominant in duplex DNA.

Although modifications at the N7 position of dG are gener-ally considered benign with regard to mutagenicity, the imid-azole ring-fragmented dG adducts are known to induce signif-icant levels of mutations in primate cells. When site-specificallymodified vectors containing an unsubstituted Fapy-dG werereplicated in either simian kidney COS7 (27, 28) or humanembryonic kidney 293T (29) cells, the adduct caused mutations,predominantly G3 T transversions and G3 A transitions, at

frequencies of �8 to 30%. Me-Fapy-dG also manifested a strongmutator phenotype in COS7 cells, inducing mutations at fre-quencies of �9 to 21%, with G3 T transversions being mostcommon (30). In contrast to this error-prone bypass observedin primate cells, both Fapy-dG (28, 31) and Me-Fapy-dG (32)were essentially non-mutagenic in E. coli.

The present study addressed the mutagenic potential andmechanisms of replication bypass of NM-Fapy-dG. Single-stranded vectors containing a site-specific NM-Fapy-dG werereplicated in COS7 and E. coli cells, and frequencies and spectraof mutations were assessed. Furthermore, the effect of NM-Fapy-dG on DNA synthesis was investigated using Saccharo-myces cerevisiae pol �, a high-fidelity DNA polymerase that rep-resents one of the two major nuclear polymerases in eukaryoticcells. The 8-oxo-dG adduct was tested under identical condi-tions as a reference lesion. In addition, differential effects onreplication were evaluated for the two anomeric forms ofNM-Fapy-dG.

Results

The general design of mutagenesis assays

Prior investigations have demonstrated that in mammaliancells, replication of single-stranded shuttle vectors containingeither Fapy-dG (27–29) or Me-Fapy-dG (30) could result intargeted mutations in up to �30% replication events, with fre-quencies and spectra being modulated by the local sequencecontext. In the case of Me-Fapy-dG, a subset of misincorpora-tions was hypothesized to be directed by the 5�-neighboringnucleotide, as the result of primer/template misalignment, fol-

Figure 1. Structures of nitrogen mustards, DNA adducts, and relevant base pairs. A, nitrogen mustards. B, the NM-induced N7-dG adduct. The chloride ofthe adduct can partially or fully hydrolyze to an alcohol (X � Cl or OH) (7). C and D, the NM-induced Fapy-dG adduct in the �- (C) or �- (D) anomeric form. E–G,structures of base pairs that were observed for 8-oxo-dG (E and F) or proposed for NM-Fapy-dG (G).

Mutagenesis of nitrogen mustard Fapy-dG

J. Biol. Chem. (2017) 292(46) 18790 –18799 18791

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 3: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

lowed by primer realignment and extension. Considering thestructural similarity of various imidazole ring-fragmented dGadducts, the current study investigated the mutagenic proper-ties of NM-Fapy-dG and in particular, addressed the potentialeffect on mutagenesis of the 5�-nucleotide. The adduct in eithera 5�-CXG-3� or 5�-TXG-3� sequence context was incorporatedinto a single-stranded pSBL vector as illustrated in supplemen-tal Fig. S1. Identically designed vectors were also engineeredto contain 8-oxo-dG. This adduct was selected to serve as areference lesion because the NM-Fapy-dG and 8-oxo-dGadducts have a certain structural similarity and in particular,both retain the canonical Watson-Crick hydrogen-bondingface of dG.

NM-Fapy-dG-induced mutagenesis in COS7 cells

The pSBL vectors containing 8-oxo-dG or NM-Fapy-dG ineither 5�-CXG-3� or 5�-TXG-3� sequence context were repli-cated in COS7 cells. Following isolation of the progeny DNAs,individual clones were obtained by transforming E. coli DH5�cells and analyzed for mutations using a combination of differ-ential hybridization technique and Sanger sequencing. Analy-ses of control vectors that contained dG at the target site wereperformed in parallel experiments. These results demonstratedessentially error-free replication of non-damaged vectors aswe have recently reported (28). Although our experimentalapproach was not designed to evaluate the bypass efficiency,no correlation was noticed between the yield of bacterial trans-formants and the type of DNA replicated. Thus, the NM-Fapy-dG lesion is unlikely to represent a very strong block forthe mammalian replication machinery. The data on the muta-genic properties of 8-oxo-dG and NM-Fapy-dG were collectedfrom a minimum of three independent transfections of COS7cells and are shown in Table 1.

Replication of 8-oxo-dG-containing DNAs resulted in a sim-ple spectrum of mutations, with G 3 T transversions beingmost common in both sequence contexts (�3.7% in 5�-CXG-3�and �5.0% in 5�-TXG-3�). In addition, low levels (�1.2%)of various deletions were observed only in the 5�-CXG-3�sequence. The differences in frequencies of both G3 T trans-versions and total targeted mutations were not statistically sig-nificant between the two sequence contexts. These results gen-erally agree with data previously reported (27, 29, 33) and areconsistent with established mechanisms of replication past

8-oxo-dG (34 – 43), in which the adducted base remains in thecanonical anti conformation opposite an incoming dC (Fig. 1E),but assumes the syn conformation opposite dA (Fig. 1F) to facil-itate generation of G3 T transversions. It cannot be excluded,however, that a subset of these transversions, occurring in the5�-TXG-3� sequence context could be formed via a primer/template misalignment.

Relative to 8-oxo-dG, the NM-Fapy-dG adduct caused tar-geted mutations at �2-fold higher frequencies (�11.6 versus�4.4%, p � 0.019, in 5�-CXG-3� and �10.6 versus �5.3%, p �0.032, in 5�-TXG-3�). G3T transversions were most commonin both sequence contexts (�8.3% in 5�-CXG-3� and �8.8% in5�-TXG-3�), although other base substitutions also occurred.The spectra of base substitutions were minimally affected bythe local sequence context. Frequencies of G3C transversionsin 5�-CXG-3� were slightly higher than in 5�-TXG-3� (�1.8 and�0.3%, respectively), but the difference was not statistically sig-nificant. Deletions were detected exclusively in the 5�-CXG-3�context and were at near background levels. Thus, whereas thestructure of the correctly matched pair can be easily envisionedfor NM-Fapy-dG (Fig. 1G), the nature of replication intermedi-ates leading to mutagenesis is less apparent. Conventionalprimer-template misalignment models can only explain a smallfraction of NM-Fapy-dG-induced mutations. Specifically withregard to the major mutation, a G3T transversion, the recog-nized model of mispair involving the Hoogsteen face in8-oxo-dG (Fig. 1F) cannot be applied to NM-Fapy-dG, becausethis face is affected by the NM modification (Fig. 1C).

NM-Fapy-dG-induced mutagenesis in E. coli

The pSBL vectors containing NM-Fapy-dG in the 5�-CXG-3� sequence were replicated in wild-type E. coli (ZK126,W3110�lacU169tna-2) and the isogenic mutant lacking allthree DNA damage-inducible (SOS) DNA polymerases, pol II,pol IV, and pol V (SF2018, ZK126 polB::Spcr dinB::Kanr

umuDC::Camr). After obtaining ampicillin-resistant clones,frequencies and spectra of mutations were assessed by a com-bination of differential hybridization and Sanger sequencing.The data collected from the two independent vector prepara-tions demonstrated that in both E. coli strains, replicationbypass of NM-Fapy-dG was significantly more accurate than inCOS7 cells. Specifically, no mutations were detected in 152clones analyzed for the wild-type, and only two of 175 clones

Table 1Mutations generated following replication of site-specifically modified vectors in COS7 cellsThe data in parentheses are the mean percentages of mutations with corresponding standard errors calculated from three or four independent transfections of COS cells.Because of variabilities, the mean percentages are not exactly equal to percentages calculated from combined results. The standard errors for the mean frequencies �1% arenot reported. — indicates no mutations were detected.

Target siteSequencecontext

Clonestested

G to TNo.

G to CNo.

G to ANo.

G del.No.

OthersNo.

8-Oxo-dG 5�-CXG-3� 344 12a (3.7 � 1.3) — — 2b (0.6) 2c (0.6)5�-TXG-3� 284 14d (5.0 � 0.4) 1 (0.3) — — —

NM-Fapy-dG 5�-CXG-3� 355 29 (8.3 � 0.8) 6 (1.8 � 1.0) 4 (1.2 � 0.1) 2e (0.6) 2f (0.6)5�-TXG-3� 319 29 (8.8 � 1.2) 1 (0.3) 5 (1.5 � 0.7) — 2g (0.6)

a Including CG to TT and GG to TC.b Including deletion CG.c 5�-C deleted (2 clones).d Including GG to TT.e Including GCXGG to TTGG.f 5�-C deleted and 3�-G to T.g 3�-G to T.

Mutagenesis of nitrogen mustard Fapy-dG

18792 J. Biol. Chem. (2017) 292(46) 18790 –18799

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 4: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

tested for the polymerase-deficient strain contained a muta-tion, which were a G3A transition at the 3�-dG and deletion ofthe 5�-dC. The latter observation implies that in E. coli, high-fidelity DNA polymerases, pol I and/or pol III, can bypass theNM-Fapy-dG adduct, predominantly producing the correctreplication products.

Replication bypass of the NM-Fapy-dG adduct by DNApolymerase �

To further elucidate the mechanisms of replication past theNM-Fapy-dG adduct, the ability of pol � to catalyze eithererror-free or error-prone bypass of this adduct was assessed.The reason for testing pol � was that this polymerase representsone of the two major nuclear DNA polymerases in eukary-otic cells and therefore, it encounters a DNA lesion prior torecruitment of any specialized polymerase. S. cerevisiae pol �was utilized because this enzyme is available as an exonu-clease-deficient mutant, in addition to its exonuclease-pro-ficient form, with both versions being extensively character-ized (40, 44, 45).

The initial set of reactions was conducted in the presence ofall four dNTPs under running start conditions. A 32P-labeled20-mer oligodeoxynucleotide primer was hybridized with con-trol dG, NM-Fapy-dG, and 8-oxo-dG templates with the lesionin the 5�-CXG-3� sequence context to initiate DNA synthesissix nucleotides upstream of the modified site (�6 primer) (Fig.2A). These data demonstrated that both lesions inhibited DNApolymerization, one nucleotide before the adducted site (�1site) and opposite to it (0 site) (Fig. 2B). NM-Fapy-dG appearedto be a stronger block to replication than 8-oxo-dG. In partic-

ular, at 2 nM pol � concentration, the amount of extension prod-ucts beyond 8-oxo-dG was approaching the correspondingvalue calculated for the control dG (�77 versus �93%). Underidentical conditions, a significantly lower fraction of primers(�36%) was extended beyond the NM-Fapy-dG site. Surpris-ingly, accumulation of the bypass products on NM-Fapy-dG-containing template was minimally changed when thefinal enzyme concentration was increased 7-fold to 14 nM

(�40%). The lack of linear relationships between the enzymeconcentration and the product formation was also observed forNM-Fapy-dG that was positioned in the 5�-TXG-3� sequencecontext (supplemental Fig. S2). An exonuclease-deficient pol �was similarly limited in its ability to replicate past NM-Fapy-dG(data not shown). We reasoned that under these experimentalconditions, NM-Fapy-dG existed as a mixture of at least twospecies that differentially affect the ability of pol � to carry outDNA synthesis. Relevant to this observation, prior analysesdemonstrated that NM-Fapy-dG exists as a mixture of unnat-ural �- and natural �-anomeric configurations (24). Thus, wehypothesized that although NM-Fapy-dG in its �-anomericform could be engaged in extendable base pairs (Fig. 1, C andG), the �-anomer (Fig. 1D) represented a severe block to pol�-catalyzed primer extension.

To evaluate the accuracy of nucleotide incorporation by pol �opposite NM-Fapy-dG, reactions were carried out in the pres-ence of individual dNTPs using a 20-mer �1 primer (Fig. 3A).In the presence of dCTP, primer extension dominated overdegradation on all five DNA templates tested, including controldC and NM-Fapy-dG in the 5�-TXG-3� and 5�-CXG-3�sequences, and 8-oxo-dG in the 5�-CXG-3� sequence (Fig. 3, Band C). In the presence of other dNTPs, the formation of theextension products was minimal or not observed. Consistentwith the analyses on pol �-catalyzed primer extensions in thepresence of all four dNTPs (Fig. 2), incorporations of dC wereinhibited opposite 8-oxo-dG (�64 versus �85% control, p �0.031) and even more notably opposite NM-Fapy-dG (�51 ver-sus �90% control, p � 0.0001, and �49 versus �85% control,p � 0.005, in the 5�-TXG-3� and 5�-CXG-3� context, respec-tively) (Fig. 3C).

Single-nucleotide incorporation experiments were also con-ducted using exonuclease-deficient pol � (Fig. 3D). In agree-ment with published data (40), this enzyme demonstrated a lowfidelity of nucleotide incorporation opposite 8-oxo-dG. Theprimers were extended better in the presence of dATP thandCTP (�78 and �64%, p � 0.030). This result was in contrastto more selective nucleotide incorporations observed onNM-Fapy-dG-containing templates; in both sequence con-texts, higher levels of the extension products were formed inreactions supplemented with dCTP (�49 and �51% in the5�-TXG-3� and 5�-CXG-3� context, respectively) than any otherdNTP (Fig. 3D). The difference between incorporation of dCand the next preferred nucleotide (dA in 5�-TXG-3� and dG in5�-CXG-3�) was statistically significant (p � 0.043 and 0.002,respectively). Thus, even in the absence of proofreading activ-ity, incorporation opposite NM-Fapy-dG by pol � appeared tobe accurate.

Further experiments were designed to address the extensionstep of pol �-catalyzed bypass. A series of 20-mer 0 primers

Figure 2. Replication bypass of NM-Fapy-dG by pol �. A, structure of the�6 primer/template substrates. B, pol �-catalyzed primer extensions. Reac-tions were carried out for 30 min at 37 °C in the presence of 5 nM DNA sub-strates and 100 �M each of the four dNTPs. A representative gel image isshown. The mean percentages of the bypass products with correspondingstandard deviations were calculated from three independent experiments.

Mutagenesis of nitrogen mustard Fapy-dG

J. Biol. Chem. (2017) 292(46) 18790 –18799 18793

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 5: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

were hybridized with the 5�-CXG-3� templates to place eitherdC or one of the three mismatched nucleotides opposite theadducted site (Fig. 3E), and primer extensions were performedin the presence of dGTP and dCTP. In the given sequence con-text, sequential incorporations of these nucleotides would gen-erate 22-mer products, with the initially existing mismatchesbeing potentially corrected by exonucleolytic proofreadingwith subsequent dC insertion. These data demonstrated that onboth control and NM-Fapy-dG templates, the reactions likely

followed the above scenario (Fig. 3F). Although 20-mer primershad distinct electrophoretic mobilities depending on the iden-tity of the 3�-terminal nucleotide, the 21- and 22-mer extensionproducts appeared indistinguishable. In contrast, the differencein the product mobilities was obvious on 8-oxo-dG-containingtemplate, specifically when dA was placed opposite the adduct.

The in vitro replication analyses suggested that only the�-anomer of NM-Fapy-dG could be bypassed by pol � and thatthe outcome of this reaction was likely to be generally accurate.

Figure 3. Fidelity of replication bypass of NM-Fapy-dG by pol �. A, structure of the �1 primer/template substrates. B, pol �-catalyzed single nucleotideincorporations. Reactions were carried out for 30 min at 37 °C in the presence of 2 nM pol �, 5 nM DNA substrates, and 100 �M of an individual dNTP. The controlreaction did not contain dNTPs or pol �. A representative gel image is shown. C, pol �-catalyzed nucleotide incorporation and exonucleolytic primer degra-dation in the presence of dCTP. The data represent analyses of images as exemplified in panel B. The mean percentages of the extension and degradationproducts with corresponding standard deviations were calculated from three or four independent experiments. D, single nucleotide incorporations catalyzedby exonuclease-deficient pol �. Reactions were carried out for 30 min at 37 °C in the presence of 2 nM D520V pol �, 5 nM DNA substrates, and 100 �M of anindividual dNTP. The mean percentages of the primer extension products with corresponding standard deviations were calculated from three or four inde-pendent experiments. E, structures of the 0 primer/template substrates and potential products of primer extensions in the presence of dCTP and dGTP. F, pol�-catalyzed primer extensions. Reactions were carried out for 20 min at 37 °C in the presence of 2 nM pol �, 5 nM DNA substrates, 100 �M dCTP, and 100 �M dGTP.A representative gel image is shown.

Mutagenesis of nitrogen mustard Fapy-dG

18794 J. Biol. Chem. (2017) 292(46) 18790 –18799

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 6: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

In contrast to 8-oxo-dG, NM-Fapy-dG did not show the pro-pensity to form an extendable mispair with dA in pol �-cata-lyzed reactions. Considering the presence of a significant frac-tion of the NM-Fapy-dG species that posed a severe, if notcomplete block to pol �-catalyzed replication, we did not find itfeasible to assess the bypass efficiency by Michaelis-Menten orother conventional kinetic analyses. However, it is worthemphasizing that under non-saturating conditions, theamounts of products beyond the target site were comparableon all three templates tested (Fig. 2B). In the presence of 0.08 nM

pol �, these numbers on dC-, NM-Fapy-dG-, and 8-oxo-dG-containing templates were, respectively �10, 6, and 7.5%, andin the presence of 0.4 nM pol �, respectively, �38, 11, and 19%.These differences are not dramatic, especially given the consid-eration that on the NM-Fapy-dG-containing template, a frac-tion of pol � was likely to be involved in futile attempts to bypassthe blocking form of the adduct.

Mutagenic properties of anomeric forms of NM-Fapy-dG inCOS7 cells

As discussed earlier, the origin of the major NM-Fapy-dG-induced mutation, the G3 T transversion, could not be uni-versally explained by primer/template misalignments. Thepatterns of nucleotide incorporations by pol � were also incon-sistent with this initially proposed mechanism. Another possi-bility to consider would be the non-instructive incorporation ofdA opposite the lesion, as has been well-described for an abasicsite (46 – 49). We hypothesized that NM-Fapy-dG in its unnat-ural � configuration (Fig. 1D) could mimic the abasic site struc-ture in the polymerase active site, thus facilitating the forma-tion of the G3 T transversions. To address this possibility, weemployed a strategy recently developed by us (28) for increasingthe fraction of the �-anomeric species in a mixed population ofDNA vectors. Following ligation of NM-Fapy-dG-containingvectors, aliquots of these were stored for several days at 4 °Cwith scaffold oligodeoxynucleotides being intact, thus preserv-ing the double-stranded environment. Based on data from aprior study (24), we anticipated that the equilibrium would shifttoward the �-form under these conditions. Further, the vectorswere incubated with Endo IV to eliminate the residual �-ano-mers and following degradation of scaffold oligodeoxynucle-otides, immediately transfected into COS7 cells to minimizethe possibility of re-equilibration to the initial �/�-anomericratio. Because no significant sequence-dependent differenceswere found with regard to mutagenic outcome, this analysiswas limited to the 5�-CXG-3� sequence context. The data col-lected from three independent transfections revealed a signifi-cant effect of the above manipulations (Fig. 4), with both theoverall frequency of targeted mutations and the frequencyof G3 T transversions decreased more than 2-fold (�4.9 ver-sus �11.6%, p � 0.014, and �3.8 versus 8.3%, p � 0.005, respec-tively). These results strongly suggest that the �-anomer signif-icantly contributed to NM-Fapy-dG-induced mutagenesis inCOS7 cells.

Discussion

The present study was designed to evaluate the contributionof the NM-Fapy-dG adduct to mutagenesis associated with the

clinical use of NMs. The data revealed that in primate cells,NM-Fapy-dG induced mutations at frequencies of �11–12%.These frequencies were higher than that caused by 8-oxo-dG inthe identical sequence contexts and were comparable with fre-quencies of mutations reported for the related Fapy-dG (27–29) and Me-Fapy-dG (30) adducts. Relative to 8-oxo-dG, NM-Fapy-dG manifested more complex spectra of mutations,although G 3 T transversions were most common in bothcases. Thus, NM-Fapy-dG has a potential to contribute to NM-induced genomic instability.

The mechanism of mutagenic bypass of NM-Fapy-dG wasaddressed by altering the identity of the 5�-neighboring nucle-otide, with the adduct being placed in either 5�-CXG-3� or5�-TXG-3� sequence. The effect of sequence context on spectraof NM-Fapy-dG-induced base substitutions was not apparent,and deletions were rare. Thus, we conclude that the primer/template-misaligned replication intermediates were not com-monly formed in these specific sequences.

Following enrichment of NM-Fapy-dG with the �-anomericform, frequencies of targeted mutations, particularly G 3 Ttransversions, were significantly decreased. This implicates the�-anomer as a major contributor to NM-Fapy-dG-inducedmutagenesis. Similar results have been recently reported by usfor unsubstituted Fapy-dG, with the major effects beingobserved on frequencies of both G3 A transitions and G3T transversions (28). It is important for understanding thebiology of various Fapy-dG adducts that they predominantlyexist as the �-anomers in double-stranded DNA (24 –26).However, all currently available data on mutagenesis ofFapy-dG (27–29), Me-Fapy-dG (30), and NM-Fapy-dG weregenerated using single-stranded vectors that are likely tocontain a significant portion of the �-anomeric species.Thus, the mutagenic potential of these adducts could havebeen overestimated in these studies, and consideration

Figure 4. Modulation of mutagenic properties of NM-Fapy-dG in COS7cells following enrichment with the �-anomer. The ligated vectors con-taining NM-Fapy-dG in the 5�-CXG-3� sequence context were incubated inthe presence of the scaffolding oligodeoxynucleotides and treated withEndo IV. Immediately following treatment, scaffolding oligodeoxynucle-otides were degraded, and vectors transfected into COS7 cells. The totalnumbers of clones analyzed were 355 for untreated vectors and 262 forEndo IV-treated vectors. The mean frequencies of mutations with corre-sponding standard errors were calculated from a minimum of three inde-pendent experiments using KaleidaGraph 4.1 software (Synergy Soft-ware). The p values were calculated using Student’s t test. The asteriskindicates significant difference (p � 0.05) between untreated versustreated conditions.

Mutagenesis of nitrogen mustard Fapy-dG

J. Biol. Chem. (2017) 292(46) 18790 –18799 18795

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 7: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

should be given to an alternative experimental system thatallows the adduct to be examined in a double-stranded DNAenvironment.

The hypothesis that the �-anomeric NM-Fapy-dG is not ahighly cytotoxic or miscoding lesion agrees with the data of thein vitro replication assays using pol �. This polymerase couldreplicate past the �-anomer and based on the results of singlenucleotide incorporations and primer extensions beyond theadducted site, the outcome of overall bypass reaction was accu-rate. This was in contrast to the propensity of pol � to incorpo-rate a dA opposite the 8-oxo-dG adduct and efficiently extendfrom this mismatched primer terminus. Thus, even though8-oxo-dG and NM-Fapy-dG both predominantly induceG3T transversions in primate cells, the underlying molecularmechanisms must be different. It is generally accepted thathigh-fidelity DNA polymerases, such as pol �, are involved inmutagenic bypass of 8-oxo-dG because of their abilities to formand extend a syn-8-oxo-dG:dA mispair (Fig. 1F) (34 –38, 40,41). Based on results of our study, we hypothesize that NM-Fapy-dG in its �-anomeric configuration will be bypassed bythese polymerases in a mostly error-free manner. The �-ano-mer, on the other hand, is expected to be a much stronger blockto replication, thereby necessitating the recruitment of low-fidelity DNA polymerases specialized in translesion synthesis(50). A subset of these must be able to carry out bypass of�-anomeric NM-Fapy-dG in primate cells, but at the expense ofmutations. We also hypothesize that in E. coli, where the arse-nal of DNA polymerases is more limited than in mammaliancells (50), bypass of the �-anomer is extremely poor or impos-sible. This would be consistent with the known strong potentialof the �-dG lesion to block replication in E. coli (51) and couldexplain a near background level of mutations in products ofreplication of NM-Fapy-dG-containing vectors in both strainstested in the present study.

Although the identities of DNA polymerases involved inreplication past �-anomeric NM-Fapy-dG are currently un-known, several candidates can be considered based on the dataof investigations on related adducts. Depletion of pol � fromhuman embryonic kidney 293T cells resulted in a significantreduction of targeted G 3 T transversions in progenies ofFapy-dG-containing vectors (29). In addition, insertions of dAopposite Me-Fapy-dG were observed in bypass reactions cata-lyzed by human pol � and pol � (52). However, both the abovestudies utilized DNAs with the adducts being placed in thesequence contexts that could facilitate insertions of dA via aprimer/template misalignment (5�-TXN-3� (29) and 5�-TXT-3�(52)). Furthermore, if the mechanism of bypass of the �-anomeris indeed similar to that of abasic sites, Rev1 and pol should beconsidered, because these polymerases were vital for replica-tion of abasic site-containing vectors in 293T cells (49). Regard-ing specifically the �-anomeric lesions, recent analyses revealedthat human pol �, pol �, and pol could incorporate nucleotidesopposite �-dG, with pol � manifesting the most accurate andefficient bypass capability (53). The two-subunit yeast pol (Rev3-Rev7) appeared to be completely incompetent on �-dG-containing templates at both incorporation and extension steps(53). Interestingly, our preliminary data demonstrated the abil-ity of four-subunit yeast pol (Rev3–Rev7–Pol31–Pol32) (54)

to synthesize DNA past NM-Fapy-dG, with the amount ofbypass products beyond the lesion site significantly exceedingthat observed for pol �. The latter result suggested that thisenzyme could either bypass the � form of the adduct or uponbinding, modulate the anomeric ratio. Future structural andbiochemical analyses will provide insights about mechanismsof NM-Fapy-dG-induced mutagenesis and in particular,address possible roles of individual DNA polymerases in muta-genic bypass of the �-anomer.

Experimental procedures

Oligodeoxynucleotides

Site-specifically modified oligodeoxynucleotides 5�-GCT-AGCXGGTCC-3� and 5�-GCTAGTXGGTCC-3�, where X isNM-Fapy-dG, were synthesized and purified by HPLC asreported previously (24). The corresponding oligodeoxynucle-otides containing either dG or 8-oxo-dG at the X site and allother oligodeoxynucleotides were synthesized and purified byIntegrated DNA Technologies.

Construction of the pSBL vector

The pSBL vector was created based on the pSB shuttle vector(55) by insertion of a stem-loop structure into the uniqueBamHI site (supplemental Fig. S1). The details of the constructpreparation are described in the legend to supplemental Fig. S1.The f1 origin-dependent production of single-stranded pSBLwas performed using DH12S E. coli (ElectroMAX, Invitrogen)following a standard procedure.

Construction of site-specifically modified vectors andmutagenesis assay

Insertion of site-specifically modified 12-mer oligodeoxy-nucleotides into EcoRV site in a stem region of single-strandedpSBL using uracil-containing scaffold oligodeoxynucleotides(supplemental Fig. S1), replication of vectors in host cells, andanalyses of the progeny vectors for mutations by differentialhybridization were performed as in our previous studies (28, 30,56, 57). To enrich the NM-Fapy-dG-containing vectors withthe �-anomeric adduct, the ligated vectors were purified bythree sequential washes with TE buffer (10 mM Tris-HCl (pH7.4) and 1 mM EDTA) using Amicon Ultra 100K centrifugalfilters (Merck Millipore, Ltd.), and stored for a minimum of 4days at 4 °C with scaffold oligodeoxynucleotides being intact.The vectors (1 pmol) were incubated with 20 units of E. coliendonuclease IV (New England Biolabs) at 37 °C for 1 h in abuffer provided by supplier (50 mM Tris-HCl (pH 7.9), 100 mM

NaCl, 10 mM Mg2Cl, 1 mM DTT), and subsequently, with 3units of T4 DNA polymerase and 5 units of uracil DNA glyco-sylase (both from New England Biolabs) at 37 °C for 30 min todegrade scaffold oligodeoxynucleotides. Vectors were purifiedby Bio-Spin P-6 columns (Bio-Rad), and immediately used fortransfection of COS7 cells.

The hybridization probes were 32P-labeled oligodeoxynucle-otides matching the clones that had no mutations in theinserted sequences (5�-ATGCTAGCGGGTCCATCG-3� or5�-ATGCTAGTGGGTCCATCG-3� for the 5�-CXG-3� or5�-TXG-3� sequence context, respectively) or acquired G 3 T

Mutagenesis of nitrogen mustard Fapy-dG

18796 J. Biol. Chem. (2017) 292(46) 18790 –18799

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 8: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

transversions at the target site (5�-ATGCTAGCTGGTC-CATCG-3� or 5�-ATGCTAGTTGGTCCATCG-3� for the5�-CXG-3� or 5�-TXG-3� sequence context, respectively). Toidentify additional clones that contained the insertedsequences, a probe was used that encompassed the junctionbetween the vector and insert sequences (5�-GTCCAT-CGCTGGATCCCG-3�). The identities of mutations in the lat-ter clones were determined by Sanger sequencing. The datacollected from independent transfections were subjected to theStudent’s t test statistical analyses using KaleidaGraph software(version 4.1, Synergy Software).

Preparation of DNA substrates for in vitro replication assays

In vitro replication assays used 59-mer oligodeoxynucle-otides as templates that were constructed according to the pub-lished procedure (56). Briefly, site-specifically modified 12-meroligodeoxynucleotides (5�-GCTAGCXGGTCC-3� or 5�-GCTA-GTXGGTCC-3�) were ligated with the flanking 5� 12-mer (5�-ACGGCCAGTGAG-3�) and 3� 35-mer (5�-GATCCCGTACT-AGTTGCTTCCTGCAGGCGTAATCA-3�) fragments, 500pmol each, in the presence of equimolar amounts of scaffoldoligodeoxynucleotides (5�-GAAGCAACTAGTACGGGATC-GGACCCGCTAGCCTCACTGGCCGTCGTTTTAC-3� or5�-GAAGCAACTAGTACGGGATCGGACCCACTAGCCT-CACTGGCCGTCGTTTTAC-3�) using 40 units of T4 DNAligase (New England Biolabs). The resulting 59-mer oligode-oxynucleotides were purified by gel electrophoresis. Thesequences of oligodeoxynucleotides used as primers were asfollows: 5�-GAAGCAACTAGTACGGGATC-3� (�6 primer),5�-AACTAGTACGGGATCGGACC-3� (�1 primer), 5�-ACT-AGTACGGGATCGGACCC-3� (0-C primer), 5�-ACTAGTA-CGGGATCGGACCA-3� (0-A primer), 5�-ACTAGTACGGG-ATCGGACCG-3� (0-G primer), and 5�-ACTAGTACGGGA-TCGGACCT-3� (0-T primer). Primer oligodeoxynucleotideswere radioactively labeled using [�-32P]ATP and T4 polynucle-otide kinase (New England Biolabs) and hybridized with tem-plates as described previously (56).

In vitro replication assay

Three-subunit S. cerevisiae pol � (Pol3–Pol31–Pol32) and itsexonuclease-deficient D520V variant were purified as reported(44, 45). Polymerase bypass reactions were conducted at 37 °Cin a buffer composed of 25 mM Tris-HCl (pH 7.5), 10 mM NaCl,8 mM Mg2Cl, 10% glycerol, 100 �g/ml of BSA, and 5 mM DTT.The reactions contained 5 nM DNA substrate and 100 �M

dNTPs. Concentrations of pol � and incubation times variedand are specified in the figures or figure legends. The reactionproducts were resolved by electrophoresis in a 15% denaturingpolyacrylamide gel containing 8 M urea in Tris borate-EDTAbuffer and visualized using a Storm PhosphoImager (GEHealthcare) or a Personal Molecular Imager (Bio-Rad). The gelimages were analyzed by Image Quant 5.2 software (MolecularDynamics) or the Personal Molecular Imager built-in software.The mean percentages of bypass and exonucleolytic degrada-tion products with corresponding standard deviations were cal-culated from a minimum of three independent experiments

using KaleidaGraph 4.1 software (Synergy Software). The p val-ues were calculated using Student’s t test.

Author contributions—All authors contributed to design of thestudy. C. J. R. provided site-specifically modified oligodeoxynucle-otides, I. G. M. conducted experiments, analyzed the data, and pre-pared the manuscript draft with significant intellectual input fromC. J. R. and R. S. L.

Acknowledgments—We thank Drs. Louise and Satya Prakash (Uni-versity of Texas Medical Branch) for a kind gift of the pSB vector, Dr.Peter M. Burgers (Washington University, St. Louis) for providing theyeast pol � and pol , and Dr. Steven E. Finkel (University of SouthernCalifornia) for sharing the ZK126 and SF2018 E. coli strains. We alsothank Tracy Johnson Salyard and Dr. Plamen P. Christov for synthesisof NM-Fapy-dG-modified oligodeoxynucleotides and Dr. Yan Sha forassistance in performing the mutagenesis assays.

References1. Rajski, S. R., and Williams, R. M. (1998) DNA cross-linking agents as

antitumor drugs. Chem. Rev. 98, 2723–27962. Hurley, L. H. (2002) DNA and its associated processes as targets for cancer

therapy. Nat. Rev. Cancer 2, 188 –2003. Valdez, B. C., and Andersson, B. S. (2010) Interstrand crosslink inducing

agents in pretransplant conditioning therapy for hematologic malignan-cies. Environ. Mol. Mutagen. 51, 659 – 668

4. Deans, A. J., and West, S. C. (2011) DNA interstrand crosslink repair andcancer. Nat. Rev. Cancer 11, 467– 480

5. Povirk, L. F., and Shuker, D. E. (1994) DNA damage and mutagenesisinduced by nitrogen mustards. Mutat. Res. 318, 205–226

6. Hosing, C., Munsell, M., Yazji, S., Andersson, B., Couriel, D., de Lima, M.,Donato, M., Gajewski, J., Giralt, S., Körbling, M., Martin, T., Ueno, N. T.,Champlin, R. E., and Khouri, I. F. (2002) Risk of therapy-related myelodys-plastic syndrome/acute leukemia following high-dose therapy and autol-ogous bone marrow transplantation for non-Hodgkin’s lymphoma. Ann.Oncol. 13, 450 – 459

7. Brookes, P., and Lawley, P. D. (1961) The reaction of mono- and di-func-tional alkylating agents with nucleic acids. Biochem. J. 80, 496 –503

8. Kallama, S., and Hemminki, K. (1984) Alkylation of guanosine by phos-phoramide mustard, chloromethine hydrochloride and chlorambucil.Acta Pharmacol. Toxicol. (Copenh.) 54, 214 –220

9. Kallama, S., and Hemminki, K. (1986) Stabilities of 7-alkylguanosines and7-deoxyguanosines formed by phosphoramide mustard and nitrogenmustard. Chem. Biol. Interact. 57, 85–96

10. Hemminki, K. (1987) DNA-binding products of nornitrogen mustard, ametabolite of cyclophosphamide. Chem. Biol. Interact. 61, 75– 88

11. Osborne, M. R., and Lawley, P. D. (1993) Alkylation of DNA by melphalanwith special reference to adenine derivatives and adenine-guanine cross-linking. Chem. Biol. Interact. 89, 49 – 60

12. Osborne, M. R., Wilman, D. E., and Lawley, P. D. (1995) Alkylation ofDNA by the nitrogen mustard bis(2-chloroethyl)methylamine. Chem. Res.Toxicol. 8, 316 –320

13. Mohamed, D., Mowaka, S., Thomale, J., and Linscheid, M. W. (2009)Chlorambucil-adducts in DNA analyzed at the oligonucleotide level usingHPLC-ESI MS. Chem. Res. Toxicol. 22, 1435–1446

14. Florea-Wang, D., Haapala, E., Mattinen, J., Hakala, K., Vilpo, J., and Hovi-nen, J. (2003) Reactions of N,N-bis(2-chloroethyl)-p-aminophenylbutyricacid (chlorambucil) with 2�-deoxyadenosine. Chem. Res. Toxicol. 16,403– 408

15. Florea-Wang, D., Pawlowicz, A. J., Sinkkonen, J., Kronberg, L., Vilpo, J.,and Hovinen, J. (2009) Reactions of 4-[Bis(2-chloroethyl)amino]benzen-ebutanoic acid (chlorambucil) with DNA. Chem. Biodivers 6, 1002–1013

16. Florea-Wang, D., Haapala, E., Mattinen, J., Hakala, K., Vilpo, J., and Hovi-nen, J. (2004) Reactions of N,N-bis(2-chloroethyl)-p-aminophenylbutyric

Mutagenesis of nitrogen mustard Fapy-dG

J. Biol. Chem. (2017) 292(46) 18790 –18799 18797

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 9: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

acid (chlorambucil) with 2�-deoxycytidine, 2�-deoxy-5-methylcytidine,and thymidine. Chem. Res. Toxicol. 17, 383–391

17. Haapala, E., Hakala, K., Jokipelto, E., Vilpo, J., and Hovinen, J. (2001) Re-actions of N,N-bis(2-chloroethyl)-p-aminophenylbutyric acid (chloram-bucil) with 2�-deoxyguanosine. Chem. Res. Toxicol. 14, 988 –995

18. Loeber, R. L., Michaelson-Richie, E. D., Codreanu, S. G., Liebler, D. C.,Campbell, C. R., and Tretyakova, N. Y. (2009) Proteomic analysis of DNA-protein cross-linking by antitumor nitrogen mustards. Chem. Res. Toxicol.22, 1151–1162

19. Michaelson-Richie, E. D., Ming, X., Codreanu, S. G., Loeber, R. L., Liebler,D. C., Campbell, C., and Tretyakova, N. Y. (2011) Mechlorethamine-in-duced DNA-protein cross-linking in human fibrosarcoma (HT1080) cells.J. Proteome Res. 10, 2785–2796

20. Chetsanga, C. J., Polidori, G., and Mainwaring, M. (1982) Analysis andexcision of ring-opened phosphoramide mustard-deoxyguanine adductsin DNA. Cancer Res. 42, 2616 –2621

21. Gruppi, F., Hejazi, L., Christov, P. P., Krishnamachari, S., Turesky, R. J.,and Rizzo, C. J. (2015) Characterization of nitrogen mustard formami-dopyrimidine adduct formation of bis(2-chloroethyl)ethylamine with calfthymus DNA and a human mammary cancer cell line. Chem. Res. Toxicol.28, 1850 –1860

22. Ide, H., Tedzuka, K., Shimzu, H., Kimura, Y., Purmal, A. A., Wallace, S. S.,and Kow, Y. W. (1994) �-Deoxyadenosine, a major anoxic radiolysis prod-uct of adenine in DNA, is a substrate for Escherichia coli endonuclease IV.Biochemistry 33, 7842–7847

23. Ishchenko, A. A., Ide, H., Ramotar, D., Nevinsky, G., and Saparbaev, M.(2004) �-Anomeric deoxynucleotides, anoxic products of ionizing radia-tion, are substrates for the endonuclease IV-type AP endonucleases. Bio-chemistry 43, 15210 –15216

24. Christov, P. P., Son, K. J., and Rizzo, C. J. (2014) Synthesis and character-ization of oligonucleotides containing a nitrogen mustard formamidopy-rimidine monoadduct of deoxyguanosine. Chem. Res. Toxicol. 27,1610 –1618

25. Patro, J. N., Haraguchi, K., Delaney, M. O., and Greenberg, M. M. (2004)Probing the configurations of formamidopyrimidine lesions Fapy�dA andFapy�dG in DNA using endonuclease IV. Biochemistry 43, 13397–13403

26. Christov, P. P., Banerjee, S., Stone, M. P., and Rizzo, C. J. (2010) Selectiveincision of the �-N5-methyl-formamidopyrimidine anomer by Esche-richia coli endonuclease IV. J. Nucleic Acids 10.4061/2010/850234

27. Kalam, M. A., Haraguchi, K., Chandani, S., Loechler, E. L., Moriya, M.,Greenberg, M. M., and Basu, A. K. (2006) Genetic effects of oxidative DNAdamages: comparative mutagenesis of the imidazole ring-opened forma-midopyrimidines (Fapy lesions) and 8-oxo-purines in simian kidney cells.Nucleic Acids Res. 34, 2305–2315

28. Sha, Y., Minko, I. G., Malik, C. K., Rizzo, C. J., and Lloyd, R. S. (2017)Error-prone replication bypass of the imidazole ring-opened formami-dopyrimidine deoxyguanosine adduct. Environ. Mol. Mutagen. 58,182–189

29. Pande, P., Haraguchi, K., Jiang, Y. L., Greenberg, M. M., and Basu, A. K.(2015) Unlike catalyzing error-free bypass of 8-oxodGuo, DNA polymer-ase � is responsible for a significant part of Fapy�dG-induced G 3 Tmutations in human cells. Biochemistry 54, 1859 –1862

30. Earley, L. F., Minko, I. G., Christov, P. P., Rizzo, C. J., and Lloyd, R. S. (2013)Mutagenic spectra arising from replication bypass of the 2,6-diamino-4-hydroxy-N5-methyl formamidopyrimidine adduct in primate cells. Chem.Res. Toxicol. 26, 1108 –1114

31. Patro, J. N., Wiederholt, C. J., Jiang, Y. L., Delaney, J. C., Essigmann, J. M.,and Greenberg, M. M. (2007) Studies on the replication of the ring openedformamidopyrimidine, Fapy�dG in Escherichia coli. Biochemistry 46,10202–10212

32. Minko, I. G., Earley, L. F., Larlee, K. E., Lin, Y. C., and Lloyd, R. S. (2014)Pyrosequencing: applicability for studying DNA damage-induced mu-tagenesis. Environ. Mol. Mutagen. 55, 601– 608

33. Moriya, M. (1993) Single-stranded shuttle phagemid for mutagenesisstudies in mammalian cells: 8-oxoguanine in DNA induces targetedG�C3 T�A transversions in simian kidney cells. Proc. Natl. Acad. Sci.U.S.A. 90, 1122–1126

34. Shibutani, S., Takeshita, M., and Grollman, A. P. (1991) Insertion of spe-cific bases during DNA synthesis past the oxidation-damaged base 8-oxo-dG. Nature 349, 431– 434

35. Einolf, H. J., and Guengerich, F. P. (2001) Fidelity of nucleotide insertion at8-oxo-7,8-dihydroguanine by mammalian DNA polymerase �: steady-state and pre-steady-state kinetic analysis. J. Biol. Chem. 276, 3764 –3771

36. Krahn, J. M., Beard, W. A., Miller, H., Grollman, A. P., and Wilson, S. H.(2003) Structure of DNA polymerase � with the mutagenic DNA lesion8-oxodeoxyguanine reveals structural insights into its coding potential.Structure 11, 121–127

37. Brieba, L. G., Eichman, B. F., Kokoska, R. J., Doublié, S., Kunkel, T. A., andEllenberger, T. (2004) Structural basis for the dual coding potential of 8-ox-oguanosine by a high-fidelity DNA polymerase. EMBO J. 23, 3452–3461

38. Hsu, G. W., Ober, M., Carell, T., and Beese, L. S. (2004) Error-pronereplication of oxidatively damaged DNA by a high-fidelity DNA polymer-ase. Nature 431, 217–221

39. Zang, H., Irimia, A., Choi, J. Y., Angel, K. C., Loukachevitch, L. V., Egli, M.,and Guengerich, F. P. (2006) Efficient and high fidelity incorporation ofdCTP opposite 7,8-dihydro-8-oxodeoxyguanosine by Sulfolobus solfatari-cus DNA polymerase Dpo4. J. Biol. Chem. 281, 2358 –2372

40. McCulloch, S. D., Kokoska, R. J., Garg, P., Burgers, P. M., and Kunkel, T. A.(2009) The efficiency and fidelity of 8-oxo-guanine bypass by DNA poly-merases � and �. Nucleic Acids Res. 37, 2830 –2840

41. Markkanen, E., Castrec, B., Villani, G., and Hübscher, U. (2012) A switchbetween DNA polymerases � and � promotes error-free bypass of8-oxo-G lesions. Proc. Natl. Acad. Sci. U.S.A. 109, 20401–20406

42. Patra, A., Nagy, L. D., Zhang, Q., Su, Y., Müller, L., Guengerich, F. P., andEgli, M. (2014) Kinetics, structure, and mechanism of 8-oxo-7,8-dihydro-2�-deoxyguanosine bypass by human DNA polymerase �. J. Biol. Chem.289, 16867–16882

43. Taggart, D. J., Fredrickson, S. W., Gadkari, V. V., and Suo, Z. (2014) Muta-genic potential of 8-oxo-7,8-dihydro-2�-deoxyguanosine bypass catalyzed byhuman Y-family DNA polymerases. Chem. Res. Toxicol. 27, 931–940

44. Burgers, P. M., and Gerik, K. J. (1998) Structure and processivity of twoforms of Saccharomyces cerevisiae DNA polymerase �. J. Biol. Chem. 273,19756 –19762

45. Fortune, J. M., Stith, C. M., Kissling, G. E., Burgers, P. M., and Kunkel, T. A.(2006) RPA and PCNA suppress formation of large deletion errors byyeast DNA polymerase �. Nucleic Acids Res. 34, 4335– 4341

46. Efrati, E., Tocco, G., Eritja, R., Wilson, S. H., and Goodman, M. F. (1997)Abasic translesion synthesis by DNA polymerase � violates the “A-rule”:novel types of nucleotide incorporation by human DNA polymerase � atan abasic lesion in different sequence contexts. J. Biol. Chem. 272,2559 –2569

47. Avkin, S., Adar, S., Blander, G., and Livneh, Z. (2002) Quantitative mea-surement of translesion replication in human cells: evidence for bypass ofabasic sites by a replicative DNA polymerase. Proc. Natl. Acad. Sci. U.S.A.99, 3764 –3769

48. Choi, J. Y., Lim, S., Kim, E. J., Jo, A., and Guengerich, F. P. (2010) Transle-sion synthesis across abasic lesions by human B-family and Y-family DNApolymerases �, �, �, , �, and REV1. J. Mol. Biol. 404, 34 – 44

49. Weerasooriya, S., Jasti, V. P., and Basu, A. K. (2014) Replicative bypass ofabasic site in Escherichia coli and human cells: similarities and differences.PLoS ONE 9, e107915

50. Goodman, M. F., and Woodgate, R. (2013) Translesion DNA polymerases.Cold Spring Harb. Perspect. Biol. 5, a010363

51. Amato, N. J., Zhai, Q., Navarro, D. C., Niedernhofer, L. J., and Wang, Y.(2015) In vivo detection and replication studies of �-anomeric lesions of2�-deoxyribonucleosides. Nucleic Acids Res. 43, 8314 – 8324

52. Christov, P. P., Yamanaka, K., Choi, J. Y., Takata, K., Wood, R. D., Guengerich,F. P., Lloyd, R. S., and Rizzo, C. J. (2012) Replication of the 2,6-diamino-4-hydroxy-N5-(methyl)-formamidopyrimidine (MeFapy-dGuo) adduct by eu-karyotic DNA polymerases. Chem. Res. Toxicol. 25, 1652–1661

53. Williams, N. L., Amato, N. J., and Wang, Y. (2017) Replicative bypassstudies of �-anomeric lesions of 2�-deoxyribonucleosides in vitro. Chem.Res. Toxicol. 30, 1127–1133

Mutagenesis of nitrogen mustard Fapy-dG

18798 J. Biol. Chem. (2017) 292(46) 18790 –18799

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 10: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

54. Makarova, A. V., Stodola, J. L., and Burgers, P. M. (2012) A four-subunitDNA polymerase complex containing Pol � accessory subunits is essen-tial for PCNA-mediated mutagenesis. Nucleic Acids Res. 40, 11618 –11626

55. Yoon, J. H., Prakash, L., and Prakash, S. (2009) Highly error-free role ofDNA polymerase � in the replicative bypass of UV-induced pyrimidinedimers in mouse and human cells. Proc. Natl. Acad. Sci. U.S.A. 106,18219 –18224

56. Kanuri, M., Minko, I. G., Nechev, L. V., Harris, T. M., Harris, C. M., andLloyd, R. S. (2002) Error prone translesion synthesis past �-hydroxypro-pano deoxyguanosine, the primary acrolein-derived adduct in mamma-lian cells. J. Biol. Chem. 277, 18257–18265

57. Minko, I. G., Kozekov, I. D., Kozekova, A., Harris, T. M., Rizzo, C. J., andLloyd, R. S. (2008) Mutagenic potential of DNA-peptide crosslinks medi-ated by acrolein-derived DNA adducts. Mutat. Res. 637, 161–172

Mutagenesis of nitrogen mustard Fapy-dG

J. Biol. Chem. (2017) 292(46) 18790 –18799 18799

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from

Page 11: Mutagenicpotentialofnitrogenmustard-induced ...neously measure various dG adducts of bis(2-chloroethyl)eth-ylamine (Fig. 1A). Following treatment of calf thymus DNA in vitro, the NM-Fapy-dG

Irina G. Minko, Carmelo J. Rizzo and R. Stephen Lloyd-anomerαadduct: Contribution of the non-canonical

Mutagenic potential of nitrogen mustard-induced formamidopyrimidine DNA

doi: 10.1074/jbc.M117.802520 originally published online September 28, 20172017, 292:18790-18799.J. Biol. Chem. 

  10.1074/jbc.M117.802520Access the most updated version of this article at doi:

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

Supplemental material:

  http://www.jbc.org/content/suppl/2017/09/28/M117.802520.DC1

  http://www.jbc.org/content/292/46/18790.full.html#ref-list-1

This article cites 57 references, 14 of which can be accessed free at

by guest on May 11, 2020

http://ww

w.jbc.org/

Dow

nloaded from