Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 271, No. 11, Issue of March 15, pp.

6185–6191, 1996
© 1996 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

Different Conformations and Site Selectivity of


HO22 -Co(III)-Bleomycin A2 and Co(III)-Bleomycin A2
Bound to DNA Oligomers*
(Received for publication, September 21, 1995, and in revised form, December 13, 1995)

Qunkai Mao, Patricia Fulmer, Wenbao Li, Eugene F. DeRose, and David H. Petering‡
From the Department of Chemistry, University of Wisconsin, Milwaukee, Wisconsin 53201

Conformational properties of HO2 2 -Co(III)-bleomycin can be modeled closely by Co-Blm (Xu et al., 1992a, 1992b).
A2 (Form I) and Co(III)-bleomycin (Form II) bound to Both form dioxygenated complexes, which undergo dimeriza-
DNA oligomers offering either principal cleavage site tion to reach a peroxyl species, HO2 2 -Co(III)-Blm (Form I) and
for the drug, d(GGAAGCTTCC)2 or d(AAACGTTT)2, have Co(III)-Blm (Form II) or the corresponding iron species. How-
been studied by NMR methods. Form I binds in slow ever, in contrast to HO2 2 -Fe(III)-Blm, its cobalt analog is stable
exchange to these oligomers. It retains most of its solu- in the presence of DNA.
tion nuclear Overhauser effects (NOEs) upon binding to The three-dimensional conformations of the two Co-Blm
either oligomer. Pyrimidinyl methyl protons from the structures have been determined by NMR spectroscopy (Xu et
metal domain of the drug make an NOE connection with al., 1994). Both forms display extensively folded structures in
a G5 2-amino proton on DNA. The bithiazole intercalates which the peptide linker is close-packed against the metal

Downloaded from www.jbc.org by guest, on June 14, 2010


between base pairs involving either C6 and T7 or T6 and
domain. In addition, in Form I the bithiazole moiety is folded
T7 of the two DNA molecules, according to NOE connec-
back over the metal domain. These structures also bind tightly
tions between the bithiazole protons and protons from
to DNA, offering the opportunity to examine them complexed
these bases and changes in the positions of their chem-
ical shifts. Form II also retains most of its solution NOEs with DNA as models for the related Fe-Blms (Xu et al., 1992b).
upon association with the first oligomer. However, in Recent communications have begun to reveal how metallo-
contrast to Form I it binds to DNA in fast exchange on bleomycins bind to DNA. NMR data on the interaction of Form
the NMR time scale over the temperature range of I with DNA 10-mer, d(CCAGGCCTGG)2 showed that the com-
5–35 °C and does not break the degeneracy of the DNA plex binds in slow exchange with DNA and that the bithiazole
proton chemical shifts. No intermolecular NOEs be- moiety partially intercalated into the base pair structure be-
tween Form II and the 10-mer have been detected. Like- tween C6 and C7 directly adjacent to the GpC cleavage site (Wu
wise, the major perturbation in chemical shift of the et al., 1994). Based on provisional NOE assignments, it was
histidine H2 and guanine G5 protons seen in Form I- proposed that the pyrimidine methyl group of Form I was in the
DNA adducts is absent in Form II-DNA. The association proximity of a 2-amino proton of G5 at the site. Another study
constant of Form II with d(GGAAGCTTCC)2 in 20 mM on the interaction of Zn-Blm with DNA concluded that Blm
HEPES buffer at pH 7.4 and 25 °C is 1.7 3 105 M21, and 1.0 binds in the minor groove in rapid exchange with oligomeric
mol of Form II bind per mol of 10-mer. DNA (Manderville et al., 1994; Manderville et al., 1995).
The finding that Form I and Zn-Blm bind differently to DNA
oligomers despite containing the same linker and DNA binding
Bleomycin (Blm)1 is an antitumor agent used to treat human domains suggests that metal domain properties play a key role
cancer. It is comprised of metal- and DNA-binding domains in the binding process. From NMR studies and associated mo-
linked by a peptide and a disaccharide unit (Fig. 1). Fe(II)-Blm lecular dynamic calculations, it appears that the metal do-
reacts with O2 and reductants to generate HO2 2 -Fe(III)-Blm, mains of Forms I and II Co-Blm bind Co(III) with opposite
which initiates single and double strand DNA cleavage and chiralities (Xu et al., 1994). Therefore, it was hypothesized that
base release (Burger et al., 1981; Petering et al., 1990; Sam et Forms I and II should associate differently with DNA. The
al., 1994; Stubbe and Kozarich, 1987). Efficient reaction is due, present study compares the interaction of Forms I and II with
in part, to the binding of the drug to DNA through the bithia- a DNA 10-mer containing a GpC site to test this hypothesis. It
zole and positively charged R group (Chien et al., 1977). There also examines the binding of Form I to another DNA oligomer
is definite site selectivity for the reactions of Fe-Blm with DNA bearing a GpT site to determine whether the features of bind-
to produce strand cleavage or base release. The drug prefers ing of Form I to DNA are common to different preferential
initial attack on the deoxyribose unit of the sequences 59- binding sites.
GpT-39 and 59-GpC-39 (McLean et al., 1989; Steighner and
Povirk, 1990). EXPERIMENTAL PROCEDURES
The redox chemistry of Fe-Blm in solution or bound to DNA Preparation of DNA Oligomers—Syntheses of d(GGAAGCTTCC) and
d(AAACGTTT) were accomplished using a Millipore Cyclone Plus DNA
synthesizer and nucleotide amidites from Millipore as starting materi-
* This work was supported by National Institutes of Health Grant
als. The method followed directions provided by the manufacturer ex-
22184. The costs of publication of this article were defrayed in part by
the payment of page charges. This article must therefore be hereby cept as modified below (Millipore, 1989). After synthesis of 30 mmol of
marked “advertisement” in accordance with 18 U.S.C. Section 1734 either oligomer, each was cleaved from the support and fully depro-
solely to indicate this fact. tected by incubation for 48 h in 30% ammonium hydroxide. Purification
‡ To whom correspondence should be addressed: Dept. of Chemistry, of the oligomers was done with a Millipore C8 reverse phase column
University of Wisconsin-Milwaukee, P.O. Box 413, Milwaukee, WI (250 3 4.6 mm). Solvent A contained 0.1 M sodium acetate, pH 6.5, and
53201. 5% acetonitrile. Solvent B was a mixture of 0.1 M sodium acetate and
1
The abbreviations used are: Blm, bleomycin; NOE, nuclear Over- 65% acetonitrile. The gradient began with 100% solvent A and reached
hauser effect; NOESY, NOE spectroscopy. 100% solvent B in 45 min with a flow rate of 3.0 ml/min. Finally, to

6185
6186 DNA-bound Cobalt(III)-Bleomycin

FIG. 1. Structure of bleomycin. Only the deprotonated amide ni-


trogen contributes to the ligand charge among the metal binding sites
(dotted). It is also assumed that the peroxyl group is protonated.
FIG. 2. Form I titration. One-dimensional proton NMR titration of
remove trityl groups, 6 ml of 80% acetic acid was incubated for 1 h with 2 mM DNAa with Form I in 20 mM phosphate buffer and 0.1 M NaCl, pH
oligomer. To separate the dimethoxytrityl groups from the oligomer, a 7.4, and 15 °C.
Sephadex G-15 desalting column was used with 20 mM ammonium

Downloaded from www.jbc.org by guest, on June 14, 2010


bicarbonate, pH 7.4, as the eluate buffer.
Association Constant of Co(III)-Blm A2 with DNA oligomers—Fluo-
Preparation of Form I and Form II and Their DNA Adducts—Blenox-
rescence spectra of Co(III)-Blm A2 were recorded with an SLM AMINO
ane, the clinical mixture primarily of Blm A2 and B2, was a gift of
4800C spectrofluorimeter. Excitation at 300 nm produces an intense
Bristol Myers Co. Bleomycin A2 was isolated from Blenoxane as de-
emission from the bithiazole moiety at 350 nm, which is quenched upon
scribed previously (Xu et al., 1994). Then, a one-to-one mixture of Form
interaction with DNA. Titration of Form I, the DNA 10-mer, with DNAa
I and Form II was made by reacting Blm A2 with Co(II)Cl2 with stirring
was carried out in 20 mM HEPES buffer at pH 7.4 and 25 °C. Plots of the
under aerobic conditions. The mixture was separated by high pressure
reduction in fluorescence intensity of the drug molecule as a function of
liquid chromatography using a 250 3 4.6-mm C18 reverse phase semi-
added oligomer were analyzed as described previously using the as-
preparative column (Bio-Rad) with the following gradient. Solvent A
sumption that Co(III)-Blm A2 binds to equivalent, noninteracting sites
contained water and 0.2% (v/v) trifluoroacetic acid. Solvent B was
on DNA (Chien et al., 1977). From the analysis one can determine K, the
methanol. The gradient changed linearly from 70% solvent A and 30%
association constant, and n, the number of base pairs/binding site,
solvent B to 63% solvent A and 37% solvent B within 10 min at a flow
using the following equation,
rate of 2.0 ml/min. When 200-ml fractions were collected, two peaks
appeared in the chromatogram representing Form II and Form I with 1/@Form II z DNAa# 5 1/nK @DNAo#@Form II# 1 1/n @DNAo# (Eq. 1)
characteristic retention times of 7.8 and 8.9 min, respectively. In order
to remove trifluoroacetic acid and methanol from the isolated Form I or in which [DNAo] represents the total concentration of DNA base pairs.
Form II mixtures, the samples were lyophilized several times in water.
The Form I or Form II DNA adducts were made by mixing isolated RESULTS
Co-Blm species with DNAa or DNAb. NMR Features of the Interaction of HO2 2 -Co(III)-Blm A2
NMR Spectroscopy—Proton NMR spectra were acquired at 500 MHz
(Form I) with d(GGAAGCTTCC)2 (DNAa)—A combination of
on a GE GN500 NMR spectrometer. Phase-sensitive two-dimensional
spectra were acquired using TPPI-States phase cycling (Marion et al., two-dimensional NMR techniques has been used to examine
1989). NOESY spectra were acquired in D2O with a mixing time of 250 the complexes of Form I and II bound to two DNA oligomers
ms at 15 and 25 °C (Jeener, et al., 1979; Macura et al., 1981). TOCSY incorporating either the GpC or GpT sites of cleavage (Petering
spectra were acquired in D2O at 15 °C with a mixing time of 80 ms using et al., 1990). The formation of 1:1 complexes of Form I and DNA
a WALTZ-17y mixing scheme (Bax, 1989; Braunschweiler and Ernst, was monitored by following the disappearance of DNA imino
1983). The proton sweep widths in the D2O spectra were set to 6024 Hz,
proton resonances. Both drug-DNA oligomer complexes were in
centered on the water resonance, with 2048 complex points acquired in
t2, and 256 complex t1 increments collected, corresponding to acquisi- slow exchange on the NMR time scale as illustrated by the
tion times of 340 and 42 ms in the two dimensions. The residual HOD observation of distinct resonances for free and bound DNA in
peak was suppressed with an on resonance DANTE pulse sequence the imino region of the spectrum at intermediate drug-to-DNA
applied during the second half of the 1-s recovery delay between scans. concentrations for the titration of d(GGAAGCTTCC) with Form
NOESY spectra in 90% H2O, 10% D2O were acquired with mixing times I (Fig. 2).
of 200 ms at 5 and 15 °C. The water resonance was suppressed using a
The binding of Form I lifts the 2-fold symmetry of the DNA
“jump and return” read pulse, with an 80-ms delay between pulses,
corresponding to excitation maxima of 6 3125 Hz from the carrier and thereby splits the degeneracy of its 1H NMR spectra. Fig. 3
frequency centered on the water resonance (Plateau and Gueron, 1982). illustrates this for the fingerprint region in the NOESY spec-
A 45° phase shift was added to the normal mixing pulse-phase cycling trum of the Form I-DNAa complex, showing a sequential NOE
scheme to minimize radiation damping (Driscoll et al., 1989). The pro- pathway for each DNA strand. The sequential cross-peak be-
ton sweep widths were set to 12,048 Hz, with 2048 complex points tween C6-H19 and T7-H6 protons is missing in this region,
acquired in t2, and 256 complex t1 increments collected, corresponding
suggesting that the drug has substantially perturbed this part
to acquisition times of 170 and 21 ms in the two dimensions. The delay
between scans was 1 s. Typically 96 scans were obtained per free of the structure. In addition, the sequential cross-peaks be-
induction decay in all two-dimensional experiments. The NMR data tween the G15 and T7 imino protons are missing in the H2O
were processed on a Silicon Graphics INDY workstation using Felix 2.3 NOESY spectrum, and these protons are significantly shifted
software (Biosym Technologies, Inc.). Sine-bell squared window func- upfield. These results are consistent with the intercalation of
tions shifted by 60° followed by 2-Hz exponential multiplication were the drug between C6-G15 and T7-A14 (Wüthrich, 1986). A list
applied in both dimensions, and the data were zero-filled to 2048 points
of intermolecular NOEs involving the bithiazole B5 and B59
in t1 prior to Fourier transformation. Polynomial base-line correction
was applied to the F2 and F1 dimensions to obtain flat base lines. protons in Table I confirms this interpretation by showing that
One-dimensional spectra of samples in 90% H2O, 10% D2O were also these protons interact with several protons in the C6-G15 and
acquired using jump and return water suppression. T7-A14 base pairs. Together with the upfield shifts in position
DNA-bound Cobalt(III)-Bleomycin 6187

FIG. 3. Fingerprint region of the 1H 250-ms NOESY spectrum of the 2 mM 1:1 Form I-DNAa complex in D2O, 20 mM phosphate (pH

Downloaded from www.jbc.org by guest, on June 14, 2010


7.4), and 0.1 M NaCl at 25 °C. a and b trace the sequential NOE pathways for each strand from 59 end to 39. The dashed circle indicates the missing
sequential NOE (C6-H19 to T7-H6).

TABLE I also signals that the metal domain interacts with the oligomer.
Bithiazoles and base intermolecular NOEs With the bithiazole intercalated between C6-G15 and T7-A14,
Complex NOE the metal domain-linker structure intact, and the metal do-
Form I-DNAa (10-mer) main in proximity to G5, it appears that Form I retains a
B59-C6 H5 6.97, 5.52 compact, folded structure when bound to DNA. Evidently,
B59-C6 NH2 6.97, 6.46 Form I can bind to a GpC site of cleavage without substantial
B59-T7 CH3 6.97, 1.70
reorganization of its solution structure, bringing the minor
B59-G15 NH 6.97, 12.3
B5-A14 H8 7.04, 7.52 groove cleavage site of the drug into apposition with the DNA
B5-A14 H19 7.04, 5.52 and metal domains.
B5-A14 H29, H20 7.04, 2.32/2.62 NMR Features of the Interaction of HO2 2 -Co(III)-Blm A2
B5-G15 H8 7.04, 7.44
B5-G15 H19 7.04, 5.58
(Form I) with d(AAACGTTT)2 (DNAb)—A similar picture is
Form I-DNAb (8-mer) emerging from an examination of the Form I-DNAb adduct.
B59-T6 NH 6.73, 13.25 The bithiazole B5 and B59 proton resonances are shifted to a
B59-T6 CH3 6.73, 1.36 higher field (8.11 and 7.77 ppm to 6.85 and 6.73 ppm, respec-
tively) upon binding of Form I to DNAb. Correspondingly, the
T6 and T7 imino protons are upfield shifted, consistent with
of the bithiazole B5 and B59 resonances from 8.11 and 7.77 ppm
the intercalation of the B5-B59 edge of the bithiazole between
to 7.04 and 6.97 ppm, respectively, these results confirm that
T6-A14 and T7-A13. Intermolecular NOEs between these parts
the sulfur-containing edge of the bithiazole intercalates be-
of the two molecules also support the binding of the bithiazole
tween C6-G15 and T7-A14. Similar data have been obtained
from an analysis of the Form I-DNAc adduct, leading to the to this region of DNAb (Table I). In addition, as with the Form
same interpretation (Wu et al., 1994). I-DNAa complex, the folded structure of the metal domain-
Form I itself retains almost all of the intramolecular NOEs linker remains in the DNA complex, but the B5-PCH3 NOE is
seen in the absence of DNAa at positions similar to those missing (Table II). Instead, the PCH3 protons make an NOE
previously reported (Table II) (Xu et al., 1994). Thus, in the connection with a G5 2-amino proton (Fig. 4B). Also, the imid-
DNAa adduct, extensive folding exists in the peptide linker azole H2 is highly shifted from 8.69 to 9.14 ppm as in Form
region, which generates a close-packed structure involving the I-DNAa. Clearly, Form I binds to both sites, GpC and GpT,
metal domain and linker similar to that in free Form I. How- with closely related conformations.
ever, the NOE between B5 and pyrimidinyl (P) methyl hydro- NMR Features of the Interaction of Form II, Co(III)-Blm A2
gens is missing. In the unbound form this NOE shows that the with d(GGAAGCTTCC)2—Form II binds to DNAa in faster
metal domain is folded over the DNA domain (Xu et al., 1994). exchange than Form I and does not break the 2-fold symmetry
Evidently, this interaction is lost upon intercalation of the of the DNAa NOESY spectrum (Figs. 5 and 6). In data not
bithiazole into DNA. shown, the fast exchange regions for the adduct existed over
The PCH3 protons at 2.55 ppm display NOE interactions the temperature ranges of 5–35 °C. Nevertheless, small pertur-
with an exchangeable base proton at 10.22 ppm. Although bations in chemical shifts of protons throughout both the drug
highly shifted, this proton is identified as a G5 2-amino proton and DNA structures indicate that a binding interaction does
on the basis of its NOE connections with G5 1-imino and G5 occur. The maintenance of the degenerate, uninterrupted NOE
2-amino hydrogens, which are its nearest neighbor protons pathway in Fig. 4 shows that Form II does not intercalate into
(Fig. 4A). Thus, the metal domain is closely associated with the the DNA base structure like Form I. Still, the B5 and B59
guanine base at the specific cleavage site and within a few protons are shifted upon interaction with DNAa. As with Form
angstroms of the binding site for the DNA domain. A large shift I, almost all of the NOEs seen in solution remain unperturbed
of imidazole H2 upon binding to DNA from 8.61 to 9.05 ppm in the DNA-bound structure (Table I). There is no NOE con-
6188 DNA-bound Cobalt(III)-Bleomycin
TABLE II
Drug-DNA complexes and CoBlm Intramolecular NOEs
Form I-DNAb Form I-DNAa Form II-DNAa
Complex NOEa Complex NOEa Complex NOEa
H2-VaCH3 9.14, 0.60 H2-TCH3 9.05, 1.16 H2-TCH3 8.79, 1.12
H2-Aa 9.14, 3.62 H2-VaCH3 9.05, 0.57 H2-VaCH3 8.79, 0.69
H2-Ab9 9.14, 3.05 H2-Aa 9.05, 3.65 H2-Aa 8.79, 3.31
H2-TCH3 9.14, 1.20 H2-Ab9 9.05, 3.20 H2-Hb9 8.79, 3.15
H4-Hb 7.57, 5.45 H2-ANH 8.79, 6.43
H4-Va 7.57, 1.17
H4-VaCH3 7.57, 0.60 H4-Va 7.52, 1.12 H4-Va 7.60, 1.13
Hb-G1 5.45, 5.32 H4-Hb 7.52, 5.42 H4-Hb 7.60, 5.51
H4-VaCH3 7.52, 0.57 H4-VaCH3 7.60, 0.69
Hb-G2 5.45, 4.27
Hb-G1 5.42, 5.36 Hb-G1 5.51, 5.23
Hb-G2 5.42, 4.10 Hb-G2 5.51, 4.08
VaCH3-Vb 0.60, 3.47 VaCH3-Vb 0.57, 3.41 VaCH3-Vb 0.69, 3.36
VaCH3-Vg 0.57, 3.66 VaCH3-Vg 0.69, 3.47
VaCH3-Vg 0.60, 3.70 VgCH3-Vb 0.85, 3.36
VgCH3-Vb 0.91, 3.47 VgCH3-Vb 0.87, 3.41 VNH-VgCH3 9.89, 0.85
VNH-VgCH3 8.68, 0.87 VNH-Vb 9.89, 3.36
TCH3-Ta 1.20, 4.08 VNH-Vb 8.68, 3.41
Ta-TCH3 4.11, 1.12
Ta-TCH3 4.02, 1.16 TNH-VaCH3 8.31, 0.69
TNH-VaCH3 9.23, 0.60 TNH-VaCH3 9.33, 0.57 TNH-TCH3 8.31, 1.12
TNH-Va 9.23, 1.17 TNH-Va 9.33, 1.12 TNH-Vb 8.31, 3.36

Downloaded from www.jbc.org by guest, on June 14, 2010


TNH-Vb 9.23, 3.47 TNH-Vb 9.33, 3.41 TNH-Va 8.31, 1.13
BNH-TCH3 8.51, 1.20 BNH-Ta 8.57, 4.02 BNH-Ta 8.04, 4.11
BNH-Ta 8.51, 4.08 BNH-TCH3 8.57, 1.16 BNH-TCH3 8.04, 1.12
BNH-Tb 8.57, 4.43
VNH-VgCH3 8.72, 0.91
VNH-Vb 8.72, 3.47
a
Ordered pair of chemical shifts in parts per million for the protons involving the NOE.

FIG. 5. Form II titration. One-dimensional proton NMR titration of


FIG. 4. Portions of the jump-return NOESY spectra (tm 5 200 2 mM of DNAa with Form II in D2O and 20 mM phosphate buffer, pH 7.4,
ms) of the 2 mM Form I-DNA (1:1). Spectra were recorded in H2O/ 15 °C.
D2O (90/10), 20 mM phosphate (pH 7.4), and 0.1 M NaCl. a, Form
I-DNAa, 15 °C; b, Form I-DNAb, 5 °C.
adopts a different conformation than Form I-DNAa-c, while
maintaining much, if not all, of the solution conformation of
nection between PCH3 and the DNA minor groove as is present this form of the drug.
in the Form I-DNA structures. Nor have any other intermolec- Association Constant of Co(III)-Blm A2 with DNAa—The
ular NOEs been assigned. Furthermore, the highly shifted G5 strikingly different properties of binding of Form I and Form II
2-amino proton and H2 proton resonances detected in Form to DNAa prompted the measurement of the association con-
I-DNA adducts are not seen with Form II; instead, they are stant of Form II with DNAa. Following a method used previ-
only modestly perturbed from their solution values (8.58 – 8.79 ously, Form II was titrated with DNAa and the extent of their
and 6.75– 6.85 ppm, respectively for H2 and G5 2-amino). In- interaction determined by the degree of quenching of the bi-
terestingly, the DNA protons in the fingerprint region which thiazole fluorescence upon binding to the oligomer (Chien et al.,
are most perturbed by the presence of Form II are those in the 1977). Fig. 7a shows fluorescence emission spectra of Form II in
GC base pairs at the ends of the 10-mer. Thus, the H6, H5, and the absence and presence of DNAa. The quenching that was
H19 protons of C9 are shifted downfield 0.05, 0.05, and 0.06 observed is qualitatively similar to that observed when metal-
ppm, respectively, and, correspondingly, those of C10 are free Blm interacts with DNA (Chien et al., 1977). In Fig. 7, b
shifted, 0.26, 0.29, and 0.09 ppm. Evidently, Form II-DNAa and c, a summary of the fluorescence titrations is presented,
DNA-bound Cobalt(III)-Bleomycin 6189

FIG. 6. Fingerprint region of the 250-ms NOESY spectrum of 2


mM Form II-DNAa adduct (1:1) in D2O, 20 mM phosphate (pH 7.4)
at 15 °C.

Downloaded from www.jbc.org by guest, on June 14, 2010


together with the a sample plot of data from one titration based
on Equation 1. Its linearity validates the simple binding model
described in Equation 1. From four titrations the average and
standard deviation for the association constant and number of
base pairs per binding site were 1.6 6 0.2 3 105 M21 and 10.2
6 0.6, respectively. The latter figure indicates that a maximum
of one drug molecule associated with each 10-mer. Considering
the magnitude of this equilibrium constant and the concentra-
tions used in the NMR experiments under similar conditions,
virtually all of the Form II added to DNAa during the NMR
titration study became bound to the oligomer.
DISCUSSION
Detailed structural information is beginning to appear about
metallobleomycins bound to DNA. The structure of a Zn-Blm
DNA oligomer has been published (Manderville et al., 1995).
NMR information about a Form I adduct with 10-mer DNA has
also been described (Wu et al., 1994). Although a structure for
this complex is not yet in hand, comparative NMR studies offer
the opportunity to survey how variations in drug structure and
DNA sequences affect adduct formation.
Previous studies have suggested that the DNA and metal
domains of ON-Fe(II)-Blm, OC-Fe(II)-Blm, O2-Fe(II)- and O2-
Co(II)-Blm-DNA, and Fe(III)-Blm interact with DNA (Albertini
and Garnier-Suillerot, 1984; Antholine and Petering, 1979; FIG. 7. Fluorescence titration of DNAa by Form II. a, emission
Antholine et al., 1981; Chikira et al., 1989; Fulmer and Peter- spectra of Form II and Form II in the presence of a ratio of 1.25:1.00
ing, 1994). Features of the Form I-oligonucleotide structure DNAa to Form II. b, summary of four plots of the decrease in fluores-
cence emission of Form II as a function of added DNAa. Error bars
described above confirm the intimate interaction of both drug represent ranges of values. Conditions were as follows: 20 mM HEPES,
domains with DNA. Indeed, in each of the three Form I-DNA pH 7.4, and 25 °C. c, double reciprocal plot based on Equation 1 to
structures that have been examined, it is an element of the determine K and n.
metal domain that associates with the G5 part of the preferred
GpC or GpT site of cleavage. The DNA domain also intercalates bithiazole intercalates into the base pair structure of all three
between several stacks of base pairs, involving C6-T7, T6-T7, Form I-DNA structures (Table I and Wu et al. (1994)). The
and C6-C7, so the site of the primary interaction of bithiazole is finding that the PCH3 group in the metal domain is in close
not unique but does include a member of the cleavage site on proximity to G5 in each structure supports the close association
the 39 side (Wu et al., 1994). of the metal domain with DNA (Fig. 4). In the Form I complex
Knowledge of Form I-DNA conformation is derived from with d(CCAGGCCTGG)2, the assignment of this NOE to an
several types of NMR information. The intramolecular NOEs interaction between a G4 or G5 2-amino proton and PCH3
in Table II show that Form I retains its solution conformation protons could not be distinguished (Wu et al., 1994). The pres-
for the linker and metal domain when bound to DNAa and ence of the same NOE in all three Form I-DNA complexes now
DNAb. The intermolecular NOEs as well as the upfield chem- supports the identification of this NOE as a G5 2-amino proton-
ical shifts of the bithiazole protons and protons of the base PCH3 proton interaction in each complex.
pairs involving bases at positions 6 and 7 indicate that the A synthesis of these structural determinants shows that
6190 DNA-bound Cobalt(III)-Bleomycin
Form I exists in each DNA complex as a structure in which the domain structure with its bithiazole and positively charged
metal domain and linker are folded similarly to those found in tail, which provides electrostatic stabilization for the drug-
the absence of DNA. Using the unbound structure as a model, DNA adduct. The net charge on the metal domain of Form I is
the metal domain is in close association with G5, the peroxide 11; the charge of the metal domain of Form II is either 1 or 21,
coordinated to Co occupies a position between G5 and T6 or C6, depending on the charge on the ligand occupying its sixth
and the bithiazole interacts between the base pairs at positions coordination position, probably either water or hydroxide.
6 and 7. Evidently, the metal domain as well as the DNA Therefore, the positively charged metal domains of both forms
domain interacts with DNA as previously hypothesized and should interact favorably with the negative charge of the DNA
participates in site selectivity at the guanine residue of the polymer and not be an obvious source of discrimination be-
GpC and GpT pairs (Xu et al., 1994). This conclusion is but- tween them. A possible answer to this question relates to the
tressed by the finding that the chemical shift of the histidine fact that Forms I and II may adopt different structures in
H2 proton, part of the metal domain, is markedly perturbed in solution; the first has the bithiazole folded over the metal
all of the DNA adducts. A role for the metal domain in site domain, and the second has the bithiazole tail region extended
selectivity has been previously hypothesized (Carter et al., from the folded metal domain-linker region (Xu et al., 1994).
1990; Kane and Hecht, 1994). Apparently, these conformations substantially remain in their
The Form I-DNA conformation is not unique among metal- DNA adducts. According to molecular dynamics calculations on
lobleomycins bound to DNA; a different one has been reported Form I and Form II, the ligand structure in each wraps oppo-
for Zn-Blm-DNA (Manderville et al., 1994, 1995). On the basis sitely about Co(III) to yield two chiral metal domains (Xu et al.,
of the present experiments, the binding mode of Form II to 1994). One result of such differential folding is that the two
DNAa is also distinct from that of Form I. Furthermore, Form metal domains cannot identically interact with DNA (Xu et al.,
II associates with DNA in a conformation that includes exten- 1994). This conclusion is consistent with the differential behav-
sive folding in the linker region that has not been observed for ior of the H2 and G5 2-amino protons in the two DNA adducts
and the lack of an NOE connection between the PCH3 protons

Downloaded from www.jbc.org by guest, on June 14, 2010


Zn-Blm either in solution or when bound to DNA (Akkerman et
al., 1988, Manderville et al., 1994, 1995). and a G5 2-amino proton in the Form II-DNA complex. A
Fluorescence titration experiments have shown that Form II related outcome is that the position of the bulky disaccharide is
binds to DNAa with an association constant comparable with different in the two conformations. A third consequence of the
ones measured for bleomycin and similar to or 10-fold larger different chiralities is that only in Form I can the bithiazole
than values for Cu-Blm under conditions of low ionic strength fold back upon the metal domain in Form I, placing the perox-
(Chien et al., 1977; Povirk et al., 1981; Roy et al., 1981). At 1.7 ide between them. In contrast, in Form II it is only possible to
3 105 M21, it is sufficiently large to insure that virtually all of fold the DNA domain over the metal domain such that the sixth
the detectable drug binds to DNA with the concentrations and ligand position is on an outer face of the metal domain, not
ionic strength used in the NMR experiments (Fig. 7). It is also between them, making it impossible for those two structures to
large enough to question whether the fast exchange behavior of bind similarly to DNA. Therefore, it is suggested that interca-
Form II in the NMR experiments seen in Fig. 5 results from lation by the bithiazole requires a specific metal domain con-
rapid exchange of Form II either bound to a particular site on figuration, which permits a particular folded conformational
the 10-mer or free in solution. relationship to exist between the metal and DNA domains. The
An attractive alternative is that Form II slides along the reason for this requirement needs further study.
10-mer such that drug exchanges rapidly between sites on the Acknowledgment—We appreciate the use of the spectrofluorimeter in
oligomer while bound to it. Other examples of sliding by ligands the laboratory of Dr. Sally Twining.
binding to DNA oligomers have been hypothesized based on
NMR studies in which association between ligand and DNA REFERENCES
does not break the symmetry of the DNA NMR spectrum Akkerman, M. A. J., Haasnoot, C. A. G., and Hilbers, C. W. (1988) Eur. J. Biochem.
173, 211–225
(Leupin et al., 1986; Pelton and Wemmer, 1990). In the present Albertini, J. P., and Garnier-Suillerot, A. (1984) Biochemistry 23, 47–53
study, the finding that DNA proton chemical shifts were most Antholine W. E., and Petering, D. H. (1979) Biochem. Biophys. Res. Commun. 91,
perturbed in the GC base pairs at the ends of the molecule may 528 –533
Antholine, W. E., Petering, D. H., Saryan, L. A., and Brown, C. E. (1981) Proc. Natl.
also be consistent with this interpretation (data not shown). Acad. Sci. U. S. A. 78, 7517–7520
The lack of any assignable intermolecular NOEs suggests that Braunschweiler, L., and Ernst, R. R. (1983) J. Magn. Reson. 53, 521–528
Bax, A. (1989) Methods Enzymol. 176, 151–168
sliding is very fast, thereby reducing the intensity of NOEs Burger, R. M., Peisach, J., and Horwitz, S. B. (1981) J. Biol. Chem. 256,
related to any particular drug-site interaction. Because the fast 11636 –11644
Carter, B. J., Murty, V. S., Reddy, K. S., Wang, S.-N., and Hecht, S. M. (1990)
exchange binding of Form II to DNA is qualitatively similar to J. Biol. Chem. 265, 4193– 4196
that of Zn-Blm, the latter complex may also be able to slide Chien, M., Grollman, A. P., and Horwitz, S. B. (1977) Biochemistry 16, 3641–3647
along the DNA molecule (Manderville et al., 1994, 1995). Chikira, M., Antholine, W. E., and Petering, D. H. (1989) J. Biol. Chem. 264,
21478 –21480.
The possibility that Form II can slide quickly along DNA Driscoll, P. C., Clore, M. G., Beress, L., and Gronenborn, A. M. (1989) Biochemistry
stands in contrast to the apparent static binding of O2-Co(II)- 28, 2178 –2187
Fulmer, P., and Petering, D. H. (1994) Biochemistry 33, 5319 –5327
Blm to DNA (Xu et al., 1992b). The latter permits such mole- Jeener, J., Meier, B. H., Bachmann, P., and Ernst, R. R. (1979) J. Chem. Phys. 71,
cules in close proximity along the DNA structure to exist for 474 – 492
extended periods without undergoing bimolecular oxidation Kane, S. A., and Hecht, S. M. (1994) Progress Nucleic Acid Res. Mol. Biol. 49,
313–352
reduction to produce Form I and Form II (Xu et al., 1992a). It Leupin, W., Chazin, W. J., Hyberts, S., Denny, W. A., and Wüthrich, K. (1986)
is hypothesized that O2-Co(II)-Blm is folded, possibly like Form Biochemistry 25, 5902–5910
Macura, C., Huang, Y., Suter, D., and Ernst, R. R. (1981) J. Magn. Reson. 43,
I, when bound to DNA (Chikira et al., 1989). The hypothesis 259 –281
that Form II may move along the DNA structure while bound Manderville, R. A., Ellena, J. F., and Hecht, S. M. (1994) J. Am. Chem. Soc. 116,
to it suggests that sliding by extended metallobleomycin struc- 10851–10852
Manderville, R. A., Ellena, J. F., and Hecht, S. M. (1995) J. Am. Chem. Soc. 117,
tures might play a role in the mechanism by which sites are 7891–7903
selected for DNA cleavage by the drug. Mao, Q., Fulmer, P., Otvos, J., Antholine, W., DeRose, E., and Petering, D. (1995)
FASEB J. 9, A1349
A major question arises from these findings: why does Form Marion, D., Ikura, M., Tschudin, R., and Bax, A. (1989) J. Magn. Reson. 85,
II not bind to DNA like Form I? Each contains the same DNA 393–399
DNA-bound Cobalt(III)-Bleomycin 6191
McLean, M. J., Dar, A., and Waring, M. M. (1989) J. Mol. Recognit. 1, 184 –192 Steighner, R. J., and Povirk, L. F. (1990) Proc. Natl. Acad. Sci. U. S. A. 87,
Millipore Corp. (1989) Millipore Cyclone Plus DNA Synthesizer: Operator’s Man- 8350 – 8354
ual, 7–1-7–30 Stubbe, J., and Kozarich, J. (1987) Chem. Rev. 87, 1107–1136
Pelton, J. G., and Wemmer, D. E. (1990) J. Am. Chem. Soc. 112, 1393–1399 Wu, W., Vanderwall, D. W., Stubbe, J., Kozarich, J. W., and Turner, C. J. (1994)
Petering, D. H., Byrnes, R. W., and Antholine, W. E. (1990) Chem. Biol. Interact. J. Am. Chem. Soc. 116, 10843–10844
73, 133–182 Wüthrich, K. (1986) NMR of Proteins and Nucleic Acids, p. 269, John Wiley and
Plateau, P., and Gueron, M. (1982) J. Am. Chem. Soc. 104, 7310 –7311 Sons, Inc., New York
Povirk, L. F., Hogan, M., Dattagupta, N., and Buechner, M. (1981) Biochemistry Xu, R. X., Antholine, W. E., and Petering, D. H. (1992a) J. Biol. Chem. 267,
20, 665– 671 944 –949
Roy, S. N., Orr, G. A., Brewer, C. F., and Horwitz, S. B. (1981) Cancer Res. 41, Xu, R. X., Antholine, W. E., and Petering, D. H. (1992b) J. Biol. Chem. 267,
4471– 4477 950 –955
Sam, J. W., Tang, X-J., and Peisach, J. (1994) J. Am. Chem. Soc. 116, Xu, R. X., Nettesheim, D., Otvos, J. D., and Petering, D. H. (1994) Biochemistry 33,
5250 –5256 907–916

Downloaded from www.jbc.org by guest, on June 14, 2010

You might also like