Chuma.cas.usf.edu

Structure and Function Department of Chemistry and Institute for Biomolecular Science, University of South Florida, Tampa, Florida 33620-5250 DOI 10.1002/med.10052 Abstract: Although most antibiotics do not need metal ions for their biological activities, there are anumber of antibiotics that require metal ions to function properly, such as bleomycin (BLM),streptonigrin (SN), and bacitracin. The coordinated metal ions in these antibiotics play an importantrole in maintaining proper structure and/or function of these antibiotics. Removal of the metal ionsfrom these antibiotics can cause changes in structure and/or function of these antibiotics. Similar tothe case of ‘‘metalloproteins,'' these antibiotics are dubbed ‘‘metalloantibiotics'' which are the titlesubjects of this review. Metalloantibiotics can interact with several different kinds of biomolecules,including DNA, RNA, proteins, receptors, and lipids, rendering their unique and specificbioactivities. In addition to the microbial-originated metalloantibiotics, many metalloantibioticderivatives and metal complexes of synthetic ligands also show antibacterial, antiviral, and anti-neoplastic activities which are also briefly discussed to provide a broad sense of the term‘‘metalloantibiotics.'' ß 2003 Wiley Periodicals, Inc. Med Res Rev, 23, No. 6, 697–762, 2003 Key words: albomycin; aminoglycosides; anthacycline; antibiotics; aureolic acid; bacitracin;bleomycin; cisplatin; function; gramicidin; ionophore; metalloantibiotics; quinolones; side-rophore; streptonigrin; structure; tetracycline Antibiotics can interact with a variety of biomolecules, which may result in inhibition of thebiochemical or biophysical processes associated with the biomolecules. This can be illustrated in theinteraction of the peptide antibiotic polymyxin with glycolipids which affects membrane function,1 inthe intercalation of the anthracyclines (ACs) into DNA base pairs which stops gene replication,2 in the Contract grant sponsor: American Cancer Society—Florida Division (Edward L. Cole Research grant); ContractGrant number: F94US F-3. Contract grant sponsor: University of South Florida, Research and Creative ScholarshipGrants and PYF Award.
Correspondence to: Li-June Ming, Department of Chemistry and Institute for Biomolecular Science University of South Florida,Tampa, Florida 33620-5250. E-mail: [email protected] Medicinal Research Reviews, Vol. 23, No. 6, 697 762, 2003ß 2003 Wiley Periodicals, Inc.
imbedding of the lipophilic antibiotic gramicidin3 and the insertion of the amphiphilic antibioticprotein colicin A into cell membrane4 which disturb normal ion transport and trans-membranepotential of cells, in the inhibition of transpeptidase by penicillin which affects cell wall synthesis,5and the inhibition of aminopeptidase by bestatin, amastatin, and puromycin which impairs manysignificant biochemical processes.6 While most antibiotics do not need metal ions for their biologicalactivities, there are several families of antibiotics that require metal ions to function properly. Insome cases, metal ions are bound tightly and are integral parts of the structure and function of theantibiotics. Removal of the metal ions thus results in deactivation and/or change in structure of theseantibiotics, such as bacitracin, bleomycin (BLM), streptonigrin (SN), and albomycin. In other cases,the binding of metal ions to the antibiotic molecules may engender profound chemical andbiochemical consequence, which may not significantly affect the structure of the drugs, such astetracyclines (TCs), ACs, aureolic acids, and quinolones. Similar to the case of ‘‘metalloproteins,''these families of antibiotics are thus dubbed ‘‘metalloantibiotics'' in our studies and are the titlesubjects of this review.
The term ‘‘antibiotic'' was originally coined by Selman A. Waksman and was used in the title of a book of his, Microbial Antagonisms and Antibiotic Substances published in 1945, and was defined as‘‘ . . produced by microorganisms and which possess the property of inhibiting the growth and evenof destroying other microorganisms.''7 However, many clinically useful ‘‘antibiotic drugs'' nowadaysare either synthetic or semi-synthetic, including many b-lactams, (fluoro)quinolones, and amino-glycosides. These (semi-)synthetic drugs and many synthetic metal complexes and organometalliccompounds that exhibit ‘‘antibiotic activities'' can be considered ‘‘synthetic antibiotics'' as thecounterparts of the originally defined ‘‘microbial-originated antibiotics'' from a broad sense of theterm. In this review, we focus on those nature-occurring metalloantibiotics and also briefly discuss afew synthetic metalloantibiotics to provide a broader view of the term ‘‘metalloantibiotics.'' Thestructures and anti-microbial, anti-viral, and/or anti-cancer activities of these natural and syntheticmetalloantibiotics will be discussed to provide further insight into their structure–functionrelationship.
Metal ions play a key role in the actions of synthetic and natural metalloantibiotics, and are involved in specific interactions of these antibiotics with proteins, membranes, nucleic acids, andother biomolecules. For example, the binding of Fe/Co –BLM, Fe/Cu–SN, Mg–quinolone, Mg–quinobenzoxazine, Mg–aureolic acid, and cisplatin with DNA impairs DNA function or results inDNA cleavage (Section 2); the involvement of Mg/Fe in the binding of TCs to the regulatory TetRprotein turns on the mechanism for bacterial resistance to TCs (Section 3); the binding ofmetallobacitracin to undecaisoprenyl pyrophosphate prohibits the recycling of the pyrophosphate tophosphate which in turn inhibits cell wall synthesis (Section 4); and the binding of metal ions toionophores or siderophores allows their transport through cell membrane which can cause disruptionof the potential across the membrane, enables microorganisms to acquire essential iron from theenvironment, or delivers antibiotics to foreign microorganisms (Section 5). The structural andfunctional roles of metal ions in metalloantibiotics have been further advanced in recent years fromextensive biological, biochemical, and physical studies,8 which are discussed herein to provide anoverview of this important and unique group of antibiotics.
DNA can bind many different biomolecules and synthetic compounds, including proteins, antibiotics,polyamines, and synthetic metal complexes and organometallic compounds.9 In the case of the veryspecific protein–DNA interaction, transcription is regulated to turn on or off a specific biologicalprocess. DNA is also a target for therapeutic treatment of disorders and diseases, such as cancers, viadirect ligand binding to it or binding to DNA-regulating biomolecules which in turn imparts DNA function indirectly. Several clinically used anti-cancer antibiotics, such as BLM and the ACs, areDNA-binding (and cleaving) agents. A better understanding of the structure of these antibiotics andtheir DNA complexes, and a better understanding of the relationship of structure, function, andtoxicity of these drugs can provide information for the design of more effective but less toxic drugs fortherapeutic treatments. The investigation of the interaction between DNA and synthetic compoundsor metal complexes can also further our understanding of DNA–ligand binding specificity whichwould provide clues for rational design of DNA-specific drug in the future.10 The structure andfunction of a few natural and synthetic DNA-targeting metalloantibiotics are discussed in this section.
Bleomycin (BLM, also known as Blenoxane) was first isolated as a Cu2þ-containing glyco-oligopeptide antibiotic from the culture medium of Streptomyces verticullus,11 and was later found tobe also an antiviral agent.12 It was soon found to be an anticancer agent and has ever since become oneof the most widely used anticancer drugs,13 most commonly used in treatment of testis cancer,lymphomas, and head and neck cancer, as well as the AIDS-related Kaposi's sarcoma in combinationwith cisplatin and adriamycin. However, it can cause life-threatening side effect, including lungfibrosis. BLM contains a few uncommon amino acids, such as b-aminoalanine, b-hydroxyhistidine,and methylvalerate, two sugars (gulose and mannose), a few potential metal-binding function-alities such as imidazole, pyrimidine, amido, and amino groups, and a peptidyl bithiazole chainconsidered to be the DNA recognition site (Fig. 1). Similar to many other nature products, BLM isproduced as a mixture of several analogues with BLM A2 and B2 being the most abundant.11 BLM isthe most extensively studied metalloantibiotic from several different view points, such as its metalbinding property, structural studies with a variety of spectroscopic methods, mechanistic study of itsoxidative DNA cleavage, investigation of its structure–function relationship, and its use as a non-heme model for investigation of dioxygen activation and DNA recognition/cleavage.14 There are afew BLM-like antibiotics which exhibit similar physical, structural, and biochemical characteristicsas BLM, which have been previously reviewed.15 1. DNA/RNA Binding and Cleavage The antibiotic mechanism of BLM has been proposed on the basis of the results from the better studiedFe and Co derivatives (however, it is the Fe form that is considered the active form in vivo because ofits higher abundance in the biological systems).14 In the presence of reducing agents, the metal ion inFe2þ- or Co2þ-BLM binds dioxygen and converts into an ‘‘activated form'' HOO-MIII-BLM probably Figure 1. Schematic structure of bleomycin (BLM) A2 and B2.The proposed metal-binding ligands based on spectroscopic studiesare in bold-phase.
via an superoxide-MIII-BLM intermediate. DNA cleavage by Fe–BLM is proposed to be carried outby the active O = FeV–BLM or O = FeIV–BLM species generated by O–O bond cleavage in theactivated form14,16 via oxidation at C4 0 and C2 0 –H proton abstraction from the deoxyribose of DNAimmediately following 5 0GC and 5 0GT sequences.14f,17 The damaged deoxyribose then breaks down,and cleavage of DNA strand occurs. The mechanism for DNA cleavage by Co–BLM has beensuggested to follow a similar mechanism via photo-activation.18 Recent studies indicated that thesequence GTAC is a hot spot for double-stranded DNA cleavage by Fe–BLM at site T.19 The sugarmoiety of BLM is important in determining the specificity of the cleavage since Fe–BLM anddeglycosylated Fe–BLM were reported to cleave d(CGCTAGCG)2 at different sites.20 In the presenceof H2O2, Fe3þ–BLM generates hydroxyl  OH free radical in the vicinity of DNAwhich is expected tocause DNA cleavage in vivo.21 More detailed discussion on the mechanism of DNA cleavage14 andassociated cytotoxicity22 can be found in the cited review articles.
Several studies indicate that Fe2þ–BLM can also bind and cleave RNA molecules,23 including tRNA and its precursors and rRNA.24 The cleavage occurs mainly at the junctions between double-stranded and single-stranded regions in RNA molecules,25 such as at C26 and A32 in E. colitRNA His .25d However, not all RNA molecules can be cleaved by Fe2þ–BLM, which include E. coli tRNATyr and tRNACys.25 These studies reveal that RNA cleavage by Fe2þ–BLM shows much higherselectivity as opposed to DNA cleavage that occurs at all 5 0GC and 5 0GT sites. Another significantdifference is that the rate for RNA cleavage is significantly slowed in the presence of Mg2þ andabolished at 0.5 mM (but not in DNA cleavage at even 50 mM25a), which has been attributed tostabilization of RNA structure by this metal ion.25d It is interesting to note that a DNA moleculeanalogous to the T-stem loop of yeast tRNAPhe is cleaved by Fe2þ–BLM at 5 0GT site as in the case ofnormal DNA cleavage, whereas the yeast tRNAPhe loop is cleaved at G53 after the corresponding GUsequence with a rate 16-times slower.25a Fe2þ–BLM can cleave DNA–RNA hybrids as well.23b,26However, the cleavage sites on the RNA strand are different from that of the RNA alone, and isinhibited at slightly higher Mg2þ concentrations equal or greater than 1 mM. In the meantime, theDNA strand in the hybrids is cleaved at all 5 0GT sequences and 5 0GC sequences to a less extent,similar to a regular double-stranded DNA.26b 2. Metal Binding and Coordination Chemistry BLM was originally isolated as a Cu2þ complex which has since been extensively studied.11 It hasalso been known to be an excellent ligand for binding with several different metal ions,27 includingMn2þ,28 Fe2þ /3þ,29 Co2þ /3þ,30 Ni2þ /3þ,31,32 Cu þ /2þ,33,34 Zn2þ,33 Cd2þ,35 Ga3þ,36 and Ru2þ 37 ionsas well as the radioactive 105Rh for use in radiotherapy.38 The d–d transitions of the Cu2þ complexesof BLM and analogues are detected at 600 nm with a molar absorptivity 110 M1 cm1. Theenergy of the d–d absorption is higher than those of many ‘‘type 2'' Cu2þ centers in the range of 650–750 nm, suggesting the presence of a strong ligand-field in a distorted 5- or 6-coordination sphere.39The metal coordination became clear after the structure of a Cu2þ complex of a biosyntheticintermediate of BLM was determined with crystallography.40 This intermediate contains all themetal-binding moieties, but lacks the sugars and the peptidyl bithiazole moiety. In this complex,the Cu2þ is bound to the ligand via imidazole, pyrimidine, the amines of b-aminoalamine, and theamide nitrogen of b-hydroxyhistidine.
The identification of nitrogen-containing ligands in Cu2þ , Co2þ , and Fe3þ–BLM complexes has also been achieved by means of electron spin-echo envelope spectroscopy through the detection of 14Nhyperfine coupling.41 This metal-binding mode has been considered to be conserved in Fe2þ–BLM. Aprevious observation of a perturbation on the ligand field and ligand-to-metal charge transfertransition bands in a few BLM congeners suggested the presence of an axial ligand which might beexchangeable.42 A later nuclear magnetic resonance (NMR) study of the diamagnetic CO adduct ofFe2þ–BLM suggested a similar metal binding site as previously determined, except that the amide group of a-D-mannose was considered to be involved in metal binding.43 Despite the disagreement, astructure of the metal center with five coordinated ligands in a distorted octahedral geometry hasemerged which leaves an open coordination site or an exchangeable site for oxygen binding.
3. Zn2þ and Co2þ /3þ Complexes and Their DNA Binding The diamagnetic Zn2þ–BLM complex of BLM has been utilized as a structural model for theparamagnetic Fe2þ–BLM complex owing to the difficulty in high-resolution NMR studies of theparamagnetic species. Previous 2D-NMR studies of Zn–BLM strongly suggested that the metal isbound to BLM through the secondary amine of b-aminoalanine, the amido-N and imidazole of b-hydroxy histidine, pyrimidine, and the carbamoyl group of mannose.44 This coordination chemistryof Zn–BLM has been suggested to be similar to that of the diamagnetic CO complex of Fe2þ–BLMbased on NMR studies.43 However, this metal coordination has recently been challenged by an NMRstudy of an analogous complex Zn–tallysomycin,45 in which five N-containing donors are suggested,including the primary amines of b-amino-Ala, pyrimidine, and the peptidyl amide and imidazole of b-(OH)His with the pyrimidine at the apex and an SS chirality. This study also excludes the binding ofthe carbamoyl group. Instead, the disaccharide covers the sixth binding site. This disagreement inaxial binding has also been raised in the study of HOO–Co3þ complexes of BLM and analoguesdiscussed below.
BLM forms a complex with Co2þ under anaerobic conditions at pH 6.8, which exhibits a nearly axial electron paramagnetic resonance (EPR) spectrum with g? ¼ 2.272 and g// ¼ 2.025 and shows alarge hyperfine coupling of A// ¼ 92.5 G attributed to the 59Co nucleus of I ¼ 7/2 and threesuperhyperfine-coupled lines of 13 G because of coupling with one 14N (I ¼ 1).46 The sharp EPRfeatures of this complex at 77 K with g 2 and g? > g// reflect the presence of a low-spin Co2þ centerof S ¼ 1/2 with the unpaired electron in the dz2 orbital overlapping with one N-containing ligand,since a high-spin Co2þ center of S ¼ 3/2 can only be observed at liquid He temperatures (andshowing g 2 and 4 features) because of its fast electron relaxation rates.47 The observation of alow-spin Co2þ center also concludes the presence of a strong ligand field as suggested based on theelectronic spectrum of Cu2þ–BLM above. Upon oxygen binding at 77 K, the EPR spectrum isdramatically changed to give g// ¼ 2.098 and g? ¼ 2.007 and a very small hyperfine coupling with 59Co of A// ¼ 20.2 G. The similar g values of 2 and the small coupling with the 59Co center stronglysuggest that the unpaired electron density is not located at the Co2þ center, but very possibly on thebound oxygen, i.e., a ligand-centered EPR spectrum.46a The study of Co2þ binding of the less activedeamido-BLM by means of EPR revealed that the fifth ligand is the amino group of b-amino-Ala.
Upon the introduction of DNA, the spectrum of Co2þ–BLM is not changed whereas the spectrum ofoxy-Co2þ–BLM is noticeably changed to give g// ¼ 2.106, g? ¼ 2.004, and A// ¼ 18.9 G.46 Thisspectral change indicates that the binding of BLM to DNA via the bithiazole rings should affect theorientation of the bound O2 molecule where the unpaired electron resides.
The Co2þ in Co2þ–BLM can form an activated ‘‘green species'' HOO–Co3þ–BLM and an inactive ‘‘brown species'' H2O–Co3þ–BLM upon treatment with peroxide.48 These low-spindiamagnetic Co3þ complexes of BLM and analogues have been extensively studied by means of 2D-NMR spectroscopy.49,50 The coordination chemistry of BLM has further been established from thesestudies (Fig. 2A). The overall structure is similar to that revealed in the crystallographic study of theCu2þ complex of the BLM bio-intermediate,40 despite the lack of consensus regarding the axialligands49,50 (i.e., alanyl-NH2 vs. mannose-CO–NH2 binding).
These Co3þ complexes can bind double-stranded DNA to form DNA2–Co3þ–BLM and DNA2–(HOO)Co3þ–BLM ternary complexes. Several ternary complexes have been investigated bymeans of 2D-NMR techniques, and their structures determined.51,52 A representative structure ofthe ternary complex (CGTACG)2–Co3þ –(OOH)–deglycopepleomycin is shown in Figure 2. Thecoordination chemistry of the Co3þ–deglycopepleomycin complex in the ternary complex is very


Figure 2. Top: The superimposed structures of the activated ‘green species' HOO-Co3þ-deglycopepleomycin (green ball-and-stick structure; Protein Data Bank ID 1A02) and the complex upon binding with d(CGTACG)2 (red stick structure without showingthe oligonucleotide) and (bottom) deglycopepleomycin-Co3þ (OOH)-(CGTACG)2 derived from NMR studies and moleculardynamic calculations (Protein Data Bank ID1A01.pdf).The Co3þ deglycopepleomycin complex is shown in red color, DNA in gray,Co3þ in pink, and the metal-bound peroxide in blue.
similar to that of the DNA-free complex; however, the peptidyl bithiazole tail is pointed away from themetal center (Fig. 2A). The metal coordinated moiety is sitting in the minor groove of DNA duplex,and the bithiazole rings are found to interact with DNA double helix via intercalation which exposesthe bound peroxide to the DNA. This intercalation binding mode was also observed in the binding ofmetal-free BLM to calf thymus DNA by means of low-frequency Raman spectroscopy.53 The bindingof these Co3þ complexes to DNA has brought the terminal oxygen of the Co3þ-bound peroxideclose to the 4 0-H of the scissile ribose (<3 A ˚ ), and has also resulted in several specific perturbations on the DNA structure. For example, in the case of d(CCAGTACTGG)2–(HOO)Co3þ–BLM,51a thebithiazole rings intercalate into the base pairs between T * 5 A and A6 T and the configuration of the T5–A6 riboses and the region C2 through C4 are found to deviate from B-form configuration.
The bithiazole was observed to span in the minor groove in the case of Zn2þ–BLM–DNA,54 different from the intercalation binding mode in the Co3þ–pepleomycin–DNA complex.52 Thisgroove-binding mode was suggested to be probably the initial binding of the metal –BLM complexwith DNA, prior to the more specific binding at 5 0-GC or 5 0-GT sequences (and intercalating into thebase pairs next to the sequences).52 However, intercalation has been suggested not necessarily arequired interaction for DNA cleavage in a study wherein the bithiazole terminus of Fe–BLM istethered to a porous glass bead, which shows similar efficacy in DNA cleavage as free Fe–BLMin solution.55 The importance of the lesion of one strand and the role of the bithiazole in the cleavage mechanism of double-stranded DNA have recently been very elegantly investigated. The binding ofHOO-Co3þ–BLM to a double-stranded DNA with a lesion site was studied and structure determinedwith 2D-NMR.56 This DNA has the sequence d(5 0-CCAAAG6 _ A8CTGGG)*d(5 0-CCCAG-T19ACTTTGG), in which the underlined site has a 30-phosphoglycolate lesion next to 5 0-phosphatomoiety and this ‘‘cleaved strand'' is connected to the other strand of a complementary sequence (with the corresponding lesion site occupied by an A) via two 5 0-3 0 hexaethylene glycol linkers. The metal-binding domain of Co3þ–BLM is located in the minor groove in close proximity of T19 and thebithiazole is most likely to partially intercalate between T19 and A20 according to nuclear Overhausereffect interactions (NOE, which is a function of the molecular rotational correlation time and inter-nuclear distance57). The structural model derived from this study suggests that the metal centerinteracts with G on the second strand at the 5 0 end of the cleaved site on the first strand as well as areorientation of the bithiazole rings upon cleavage of the first strand.
4. Paramagnetic Fe2þ/3þ Complexes The binding of Fe3þ to BLM at slightly alkaline conditions forms a low-spin complex of S ¼ 1/2(g ¼ 2.41, 2.18, and 1.89),58 in which the sixth position is occupied by a hydroxide based onresonance Raman spectroscopy.59 An oxy-form of Fe–BLM is formed by introducing dioxygen toFe2þ–BLM in the absence of reducing agent and DNA. This oxy form has been determined to be asuperoxide O  –Fe3þ–BLM complex based on its 57Fe Mo¨ssbauer spectrum.60 In the presence of a reducing agent, the activated hydroperoxide HOO–Fe–BLM complex is formed which is the activespecies for DNA cleavage.58b A recent theoretical study suggested that both heterolytic and hemolyticcleavage of the O–O bond in the active peroxo complex are not favorable based on energetics andreaction specificity, which concludes a direct attack on DNA by the hydroperoxide of the activecomplex.61 The Fe2þ–BLM complex is paramagnetic (S ¼ 2), which has been studied by means of NMR techniques.62,63 In paramagnetic metal complexes, the NMR signals of nuclei near the metal centercan be hyperfine-shifted outside the ‘‘regular'' spectral range (i.e., 13 ppm for 1H-NMR and200 ppm for 13C-NMR) by the unpaired electron(s) to afford a large spectral window that may reachmore than 100 ppm for 1H-NMR. In the meantime, the nuclear relaxation times are dramaticallyshortened which are proportional to the sixth power of metal-nucleus distances.64 This paramagneticcomplex exhibits many hyperfine-shifted 1H-NMR signals in a spectral window of 230 ppm (Fig. 3),which have been assigned to the protons of the coordinated ligands or the protons near the metal by theuse of 1D- and 2D-NMR techniques.63 A structural model of this metal complex has been built by theuse of the distance-dependent nuclear relaxation times as constraints. This structural model turns outto be similar to the structural models built on the basis of the Co3þ–BLM complexes and their ternarycomplex with the oligonucleotide duplex discussed above. The Fe2þ complex of a BLM congenerpeplomycin and its derivatives have been studied with X-ray absorption spectroscopy which revealsthat an axial ligand may affect the Fe(II) dp ! pyrimidine back-bonding as previously observed42which may stabilize the superoxide intermediate, consistent with the auto-oxidation rate of the metalcenter in the complexes.65 A recent 1H-NMR study of the paramagnetic Co2þ–BLM complex at pH 6.566 corroborates the metal coordination chemistry obtained in the Fe2þ–BLM study.63 A possible involvement ofmannose-amido group as the sixth ligand was proposed. Co2þ–BLM at slightly higher pH of 6.8 waspreviously determined by means of EPR to have a low-spin Co2þ center,46 which would exhibit onlybroad hyperfine-shifted 1H-NMR features.64 The observation of sharp 1NMR features in Co2þ–BLM Figure 3. Hyperfine-shifted1H-NMR spectrum of high-spin Fe2 þ BLMin D2Oat pH meter readingof 6.5.The shifted signals outsidethe regular 0 10 ppm window are because of protons in close proximity of the paramagnetic Fe2þ ion.
reflects the presence of a fast-relaxing high-spin Co2þ center. The NMR and EPR studies suggest apossible presence of a high-spin to low-spin transition that is controlled by pH. However, thishypothesis cannot be verified because of the lack of cross investigations of the complex at lower pHwith EPR at liquid He temperatures and the complex at higher pH with NMR in these studies.
5. Synthetic Analogues and Biosynthesis A number of BLM-analogous compounds have been synthesized that contain the metal bindingmoieties of BLM.67–69 These synthetic analogues form complexes with several metal ions, includingCu2þ and Fe2þ/3þ , and are able to cleave DNA molecules similar to the cleavage pattern by BLMcomplexes. A recent revisit of the Fe–BLM mimicking complex Fe–PMAH67c (PMAH ¼ 2-[N-(aminoethyl)amino)methyl]-4-[N-[2-(4-imidazolyl)ethyl]carbamoyl]-5-bromopyrimidine) showedthat HOO–Fe3þ–PMA (with low-spin features of g ¼ 2.22, 2.17, and 1.94 and a noticeable high-spin feature at g ¼ 4.3) can be formed by reacting Fe3þ–PMA with H2O2, but not with iodosyl-benzene and a base that was previously reported.67c This result is consistent with the observation in anearly study of Fe3þ–BLM.70 Several lipophilic ligands analogous to the metal-binding moiety of BLM have been synthesized that comprise a 4-alkoxypyridine with methylhistamine or methylethylenediamine moiety and a longhydrocarbon chain in the alkoxy moiety for the lipophilicity.71 These ligands bind Cu2þ and formmicelles with critical micelle concentration in the range of 0.9–1.4  104 M. The catalyticproperties of these complexes were not tested in this study. Some pyridine-containing BLM analogueswere synthesized, and investigated with spectroscopic and crystallographic techniques.72 The lmax of650 nm of Cu2þ complexes is significantly longer than that of Cu2þ–BLM11 which indicates aweaker ligand field in the complexes of these analogues, whereas the g values of 2.21–2.22 and 2.04–2.05 of these complexes are close to those of Cu2þ–BLM and its analogues (2.21–2.25 and 2.06)which suggests the presence of a similar axially symmetric magnetic environment of the Cu2þ centerin these complexes.27 BLM is a natural peptide–ketide hybrid (Fig. 1), like the cyanobacterial hepatotoxins such as cylindrospermopsin. The biosyntheses of many peptides and polyketides and their hybrid conjugates(including a number of antibiotics such as BLM, ACs, bacitracin, and some ionophore antibiotics)follow a nonribosomal pathway catalyzed by large clusters of peptide and ketide synthases/synthetases and peptide/ketide ‘‘hybrid'' synthetases, respectively.73,74 The genes of the synthases/synthetases of peptides and polyketides from microorganisms have recently been analyzed andcloned and the enzymes further studied,73–75 including those of BLM and bacitracin (Section 4).
BLM has been verified to be produced by synthetase clusters comprised of polyketide synthase andpeptide synthetase modules.76 These peptide and polyketide synthetases are comprised of a multi-domain modular structure for the catalysis of the initiation of the synthesis via ATP-activatingformation of thioester linkage to the enzyme, elongation mediated by condensation of the thioester-linked amino acid and/or peptide on the peptide carrier domain following a mechanism not yet fullyunderstood, and termination of the peptide or polyketide chain by a thioesterase domain via transfer ofthe final product to a serine in the thioesterase followed by hydrolysis.77 The reactant amino acids orcarboxylates are specifically recognized and covalently linked to the different domains beforetransferred to an intermediate peptide or polyketide chain. Changing of the stereochemistry is carriedout by epimerization domains in the enzyme complex. The studies of several peptide and polyketidesynthetases and their hybrids, including crystallographic studies of the adenylation domain and anNMR study of a peptidyl carrier domain,78 have greatly enhanced our understanding of the structureand mechanism of this superfamily of ‘‘mega enzymes.'' Since these synthetase complexes possessenzymatic activities toward the syntheses of secondary metabolites,75 thus are potential targets fordrug discovery in the production of potential bio-active peptides and polyketide as well as theirhybrids.79 B. Aureolic Acids The glyco-antibiotic aureolic acid family produced by Streptomyces species is comprised of severalmembers with similar structures, including chromomycin A3 (ChrA3), mithramycin (Mit, producedby S. plicatus and also known as plicamycin), olivomycin, and variamycin, which exhibit activitiestoward Gram-positive bacteria, DNA viruses, and tumors.80 However, high toxicity has limited theiruse as clinical antibiotics and anti-tumor agents. Mithramycin has been tested against severalmalignant diseases since its discovery,81 and has a limited use for the treatment of the Paget'sdisease82 and for the treatment of hypercalcemia83 (however, controversies have also raised84). BothChrA3 and Mit have recently been found to be potent inhibitors of neuronal apoptosis induced byoxidative stress and DNA damage in cortical neurons.85 Thus, these antibiotics may be effectiveagents for the treatment of apoptosis-associated neurological diseases, which suggests that sequence-selective DNA-binding drugs may serve as potential neurological therapeutics.
1. Structure of Aureolic Acids These antibiotics contain a metal-binding b-ketophenol chromophore, a highly functionalizedaliphatic side chain, and a disaccharide and a trisaccharide chains important for DNA binding andinhibition of DNA transcription (Fig. 4). The identity of the sugar chains and the sequence of the sugarlinkage of this drug family were first established by partial hydrolysis of olivomycin A86 and Mit.87The structure of ChrA3 has later been further studied by means of 1H- and 13C-NMR spectroscopy.88The structure of Mit has also been studied by the use of synthetic and NMR techniques,89 and hasrecently been fully determined by means of 2D-1H and 13C homonuclear and heteronuclear methodswhich is shown in Figure 4.90 The structures of the DNA complexes of the drugs have also beeninvestigated by the use of 2D-NMR techniques in recent years, which are discussed in a later section.
2. Role of Metal Ions in the Action of Aureolic Acids A divalent metal ion, such as Mg2þ, Co2þ, Zn2þ, or Mn2þ, is required for aureolic acid to bind to adouble helical DNA to form a drug2–metal–(DNA)2 ternary complex.91–93 The metal–drug2complex in the ternary complex is bound to DNA in the minor groove with a high preference to GCsites and a length of approximately six base pairs based on 1H-NMR,94 DNA footprinting,95 andbiochemical96 studies. Consequently, these antibiotics can inhibit transcription of the genes that haveG-C-rich promoter sequences,96 such as the c-myc proto-oncogene97 (this binding might be non-specific98) that regulate cell proliferation and also controls the expression of b-galactosidase.97,99These studies led to further investigations of the interaction between this antibiotic family and doublehelical DNA of different sizes and sequences.100 In addition to the above metal ions, Mit has been Figure 4. A schematic structure of mithramycin. The metal binding b-ketophenol moiety is shown with thick lines. All drugs in theaureolic acid family have the similar metal binding moiety, but vary in the sugar chains which have been suggested to cause differentinteractions with DNA.
determined to bind several other metal ions, including Ca2þ, Cd2þ, Tb3þ, Gd3þ, and alkali metals.101Metal:drug4 complexes are suggested to form for Ca2þ, Tb3þ, and Gd3þ; however, which does notassist the binding of the drug molecules to DNA.
Some early studies of Mg2þ binding of Mit showed that two different complexes can be formed which exhibit different absorption and CD spectra, in which the better known 1:2 Mg2þ–Mit2complex formed at low Mg2þ concentrations and a 1:1 Mg2þ–Mit complex formed at higher Mg2þconcentrations.93 Interactions of these two complexes with bulk DNA, polynucleotide, and anoctanucleotide are observed to be different. For example, while the 1:1 complex interact with theB-DNA-representing poly(GC)  poly(CG) and the A-DNA-representing poly-G  poly-C in a similarfashion, the 1:2 complex shows distinct interaction patterns with the two DNA forms. Since cellularMg2þ concentration varies significantly in neoplastic tissues,102 these two complexes can be expectedto form to certain extents and are considered important for in vivo action of the drug.93e 3. Role of Sugars The importance of the sugar moiety in antibiotic activity of this family has been established in earlystudy of the different congeners and derivatives of this antibiotic family.103 Removal of the sugarmoiety E (Fig. 4) from olivomycin A (affording olivomycin D) and ChrA3 (affording ChrA4) results insignificant loss of antibiotic activity. Moreover, the derivatives with only one sugar and the aglycones(without any sugar) are inactive. The involvement of the sugar chains in stabilization of the 1:2complexes M2þ–(ChrA3)2 (M ¼ Mg and Ni) in methanol solution has been suggested.104 On thecontrary, the deglycosylated chromomycinone forms 1:1 complexes with these two metal ions. Inaddition, Ca2þ was determined to form only a 1:1 complex with ChrA3 in methanol as opposed to aprevious observation in aqueous solution.101 The metal complexes of ChrA3 and Mit are found to interact differently with A- and B- representing DNA sequences. Since these two congeners differ in the sugar moieties (the sugars ofChrA3 are acetylated), this observation indicates the significance of the sugar chains in the interactionof these antibiotics with DNA.93 The difference between ChrA3 and Mit has also been shown in theirbinding with the oligonucleotide d(ACCGGGT)2, wherein ChrA3 forms a drug2–Mg2þ–(DNA)2ternary complex whereas Mit has been proposed to afford a (drug2–Mg2þ )2–(DNA)2 ternarycomplex based on their NMR spectra.105 This difference has also been attributed to the differencein the sugar chains in these two congeners. The role of the sugar chains has further been investigat-ed with a simple synthetic analogue, in which a simple triethylene glycol chain is attached to ab-ketophenolate aromatic ring structure as the aglycone of ChrA3.106 This simple model forms M2þ–ligand2 complexes (M ¼ Co and Mg) similar to ChrA3. Preliminary study by these authors shows thatthis complex can bind DNA. The above studies further corroborate the significance of the sugar chainsin metal binding and in the binding of this drug family to double-stranded DNA.
Total synthesis of aureolic acid has been attempted,107 wherein stereoselective syntheses of aryl 2-deoxy-b-glycosides and the A-B disaccharide of olivomycin have been achieved. Since manyantibiotics are found to be glycosylated, such as BLM, aureolic acid, aminoglycoside, and ACfamilies discussed in this review, further exploration of glycosylated metal complexes and their DNAbinding properties should be encouraged.
4. (Aureolic Acid)2–Mg2þ–(DNA)2 Ternary Complexes The requirement of divalent metal ions for the binding of aureolic acid to DNA has been fullyestablished by means of 1H- and 31P-NMR techniques, in which the spectral features of DNA arechanged upon the binding of the drug in the presence of Mg2þ.94 Several palindromicoligonucleotides have been used for NMR studies, wherein the addition of 1:2 Mg2þ–(Mit)2 toDNA afford ternary complexes whose 1H- and 31P-NMR spectra are completely different from thoseof the parent DNA molecules. The interactions between the drug and DNA and between the two bound


Figure 5. Stereo view of the structure of the ternary complex Mit2 Mg2þ (TCGCGA)2 obtained with 2D-NMR techniques (ProteinData Bank ID 146D.pdb). The complex has a 2-fold symmetry with the two drug molecules residing in the minor groove of the DNAduplex.
drug molecules in the drug2–Mg2þ–(DNA)2 ternary complexes have been revealed with 2D-NMRtechniques, from which the structures have been built as illustrated in Figure 5.
When a DNA sequence contains two GC sites separated by a few base pairs, such as the decanucleotide (TAGCTAGCTA)2, binding of two equivalents of (drug)2–Mg2þ complex to theDNA becomes possible.108 The introduction of two Mit drug molecules and one Mg2þ to thisdecanucleotide forms a complex with half of the DNA molecule bound with the drug complex, i.e., the(TAGCTA . . )2 moiety on one end of the helix is bound with the drug complex whereas the samemoiety on the other end is not. This binding mode breaks the symmetry of the palindromic DNAsequence, which results in doubling the number of NMR signals. Upon the addition of anotherequivalent of Mit2–Mg2þ recovers the palindromic symmetry of the complex. The structure of thisunique ternary complex (Mit2–Mg2þ)2–(TAGCTAGCTA)2 has been determined by the use of 2D-NMR techniques and molecular dynamic calculations, which can be retrieved from the Protein DataBank (PDB ID 207D.pdb). The above studies laid a good foundation for future studies of aureolic acidbinding to DNA of different lengths and sequences.
5. Paramagnetic (Aureolic Acid)2–Co2þ–(DNA)2 Terminal Complexes The diamagnetic Mg2þ can be replaced with a paramagnetic Co2þ for the binding of aureolic acid toDNA duplex to afford a ternary drug2–Co2þ–(DNA)2 complex.91a Because of the paramagnetism ofCo2þ, protons near the metal center are hyperfine-shifted64 to afford a 1H-NMR spectrum with a widespectral window of 100 ppm as represented in Figure 6 for the complex Mit2–Co2þ–(ATGCAT)2.91a There are more than 50 signals well resolved in the large spectral window, Figure 6. Hyperfine-shifted 1H-NMR spectrum (360 MHz) of the ternary complex Mit2 Co2þ (ATGCAT)2 at pD 8.0 obtained at40C.
representing a rare ‘‘high resolution'' 1H-NMR spectrum of a paramagnetic species. The good signalresolution allows further extensive study of this complex. The ternary complexes (ChrA3)2–Co2þ–(TTGGCCAA)2 and other complexes of longer oligonucleotides109 allow nuclear Overhauser effect(NOE)57 to be clearly detected for better signal assignment. However, information about through-bond nuclear interaction cannot be obtained because of the large signal widths of the hyperfine-shiftedsignals attributable to large molecular size of the ternary complex. Nevertheless, combining thedistance constrains derived from nuclear relaxation times and the geometry-related dipolar shift,64 astructure of the ternary complex has been constructed (Protein Data Bank ID 1EKH.pdb and1EKI.pdb)109 which is similar to the structures derived from the previous 2D-NMR studies of thediamagnetic Mg2þ complexes of this antibiotic family discussed above.
Streptonigrin (SN, also known as rufochromomycin and bruneomycin) is a metal-binding quinone-containing antibiotic produced by Streptomyces flocculus110 (Fig. 7). This antibiotic has been shownto inhibit several tumors and cancers (e.g., lymphoma, melanoma, and breast and cervix cancers) aswell as viruses in some early in vitro and clinical observations.111,112 While SN is active towardmammalian cells at the chromosome level, it is found to be much less effective against insect celllines.113 A recent study shows that SN also exhibits ionizing radiation-like damage toward Ataxistelangiectasia heterozygote cells.114 Despite the potency of SN, high toxicity and serious side effectsof this antibiotic have reduced its clinical value, and limit its use only as an experimental anti-tumoragent.111,112 Nevertheless, because of its anti-tumor potency and unique structure, SN has served as alead drug molecule for chemical modification and synthesis of new compounds to correlate thestructure features with the biological activity and toxicity of this potent antibiotic.115 1. Action of Metallo-SN SN is known to bind different transition metal ions to function properly.116,117 The interaction ofmetal –SN complexes with DNA has been proposed on the basis of some optical studies.118 A redoxactive metal ion such as Fe and Cu is required for this antibiotic to exhibit full antibiotic and anti-tumor activities.119,120 The redox-active Fe and Cu complexes have been shown to accelerate SN-mediated DNA scission in the presence of NADH, thus enhance the anti-tumor activity of thisantibiotic.121–123 These results indicate that metal ions are possibly directly involved in the action ofSN. However, the metal binding mode and structure of these metal complexes could not be definitelydetermined in these studies. Particularly, two different configurations of the drugs are possible formetal binding (Fig. 7) with the metal bound through either the quinolinequinone-aminefunctionalities based on the crystal structure124 or the quinolinequinone-picolinate functionalitiesthat requires a significant twist of the crystal structure.
Figure 7. Schematic structures of streptonigrin (SN). The structure A is metal-free drug determined by means of crystallography,whereas the structure B represents the configuration upon metal binding as determined by means of NMR relaxation.The formationof structure B requires a dramatic twist of the C2 C2 0 bond in structure A.
Since SN contains a quinone moiety, it may share some common mechanistic characteristics with other quinone-containing antibiotics125 such as the ACs (discussed in Section 2.D ‘‘Anthracyclines'')in terms of in vitro and in vivo DNA and RNA cleavage and inhibition of cancer growth via inter-ference with cell respiration and disruption of cell replication and transcriptional control.116,119,126The metal –SN complexes can be reduced to their semiquinone forms by NADH, which then caninduce cleavage of DNA. This process is inhibited by superoxide dismutase and catalase, indicatingthe involvement of superoxide and peroxide.119,121 Reduction of this antibiotic in the presence of abound metal ion is also confirmed by the detection of EPR signals attributable to the reducedsemiquinone form.127 Metal chelators and an antioxidant are found to prevent SN-induced DNAdamage and cytotoxicity,128 which supports the involvement of metal ions in the action of SN.
2. Metal Complexes of SN Zn2þ binds SN to afford a few different complexes with different metal binding modes at varioustemperatures, in which a 1:1 metal –drug complex is the predominant complex.129 A recent study ofthe crystal structure of a Zn2þ complex that mimics the metal-binding moiety of SN showed thebinding of the metal to the quinolinequinone-picolinate functionalities,130 corroborating the struc-tures of several paramagnetic metal complexes of the drug determined by means of NMR techniquesdiscussed below. The interaction of Zn2þ–SN with DNA and oligonucleotides has been investigatedwith 1H- and 31P-NMR spectroscopy. This study concluded the requirement of metal ion for SNbinding to DNA131 and revealed sequence preference in DNA binding of this antibiotic, in which thebinding of Zn2þ–SN to d(GCATGC)2 shows noticeable spectral changes whereas the complex doesnot affect the spectra of d(ATGCAT)2.
SN can bind several different paramagnetic metal ions, including Co2þ, Fe2þ, and Yb3þ ions, with large formation constants to form 1:1 metal –SN complexes.132 The paramagnetic Fe2þ, Co2þ,and Yb3þ complexes of SN have been studied with 1H-NMR spectroscopy and relaxation, and theirstructures have been determined.132 The study of Fe2þ–SN complex is particularly important since itis considered an active form of this drug that exhibits enhanced activity toward DNA destruction bothin vitro and in vivo.122 The hyperfine-shifted 1H-NMR signals of these paramagnetic complexes havebeen fully assigned. The proton-metal distances derived from the relaxation times of the hyperfine-shifted signals in these complexes match those of the complex with the metal located at thequinolinequinone-picolinate site (structure B, Fig. 7), but not the quinolinequinone-amine site basedon the crystal structure (structure A). This configuration requires a significant twist of the C2–C2 0bond by 180 in the crystal structure124 of the drug.
The introduction of poly[dA-dT] to reduced Cuþ–SN complex causes some small changes in chemical shift of the 1H-NMR signals of the complex (0.22–0.31 ppm), which was suggested to beattributed to the binding of this complex to the DNA duplex.123 The hyperfine-shifted 1H-NMR signalsof Co2þ–SN complex are found to be significantly changed upon addition of calf thymus DNA orpoly[dA-dT] (the chemical shifts of two hyperfine-shifted signals are shifted by 20–40 ppm),132which are also indicative of direct binding of the complex with DNA. Along with the DNA bindingstudy of Zn2þ–SN complex, these studies indicate the significance of metal ions in the action of thisantibiotic.
D. Anthracyclines Anthracycline (AC) antibiotics133 are produced by Streptomyces species. Soon after their discovery,they were found to exhibit a wide spectrum of antineoplastic activity toward both solid andhematologic tumors and cancers.134 In addition, an AC antibiotic has recently be found to exhibitantifungal activity.135 Despite their severe cardiotoxicity136 (e.g., cardiomyopathy) and otherside effects,137 these antibiotics have been widely used as dose-limited chemotherapeutic agents forthe treatment of human cancers such as acute leukemia. The side effects have been attributed to the toxicity of these drugs toward mitochondria,138 leading to disturbance of bioenergetics, inhibitions ofenzymes, oxidation of lipids, disorders of membrane, and oxidative stress. The less toxic adriamycin(doxorubicin) has currently been widely prescribed as a chemotherapeutic agent in association withother antineoplastic agents, such as BLM and cisplatin. In addition, new AC antibiotics and theirchemical derivatives are still found or synthesized,139,140 which may provide potential clinical use inthe future.
The antineoplastic activity of AC antibiotics has been mainly attributed to their strong interactions with DNA in the target cells. The AC family members possess a quinone-containingchromophore and an aminoglycoside side chain.133 The structures of the representing members of thisfamily daunomycin (daunorubicin) and adriamycin are shown in Figure 8.141 There are a fewmembers of the AC family that contain more extensive sugar chains, such as b-rhodomycin contains amonosaccharide and a trisaccharide, cinerubins, marcellomycin, and rhodirubins have a trisac-charide, and musettamycin has a disaccharide chain.133 The redox activity of the AC ring plays a keyrole in the action of these drugs. In addition, the metal ion bound to the 11,12-b-ketophenolate site isalso thought to be involved in some actions of these antibiotics.
1. Action of AC and Metal –AC Complexes The action of this drug family has been considered to be attributable to their redox activity and DNA-binding capability.133,142,143 Two pathways have been proposed for these drugs to deform DNAstructure and terminate biological function of DNA:2 (a) intercalation of the drugs into the base pairswith the sugar chain sitting in the DNA minor grooves which involves hydrogen bonding,electrostatic, van der Waals, and hydrophobic interactions (a representing AC–DNA structure isshown in Fig. 9); and (b) a free radical damage of the ribose. The intercalation of AC drugs to DNAdramatically distort the DNA structure which thus prohibits transcription. A number of crystal144 andNMR145 structures of different AC–DNA complexes have been resolved. A representing AC–DNAstructure nogalamycin2-d(TGTACA)2 is shown in Figure 9.144h These structural studies allowdetailed comparison of sequence specificity of the drug binding and the different modes for thebinding of different drugs with DNA.
These antibiotics can be reduced to their semiquinone forms by biological reducing agents, such as NADH and NADPH. Superoxide anion radical (O) and H 2O2 can be produced from dioxygen upon receiving electrons from the semiquinone. Then, hydroxyl radicals can be generated, which canattack cell components, such as membrane and DNA, and impair cell functioning. In the presence ofascorbic acid and H2O2, hydroxyl radicals can also be generated by Cu2þ and Fe3þ–adriamycin.146The radicals generated during the redox cycle of ACs and their Fe complexes have been considered thecause of the cardiotoxicity.147 However, a recent study showed that the capability of producing freeradicals of ACs is not directly related to their cardiotoxicity. For example, although the 13-hydroxyderivatives of ACs are more cardiotoxic, they are less effective producers of oxygen-based freeradicals.148 The 13-hydroxy metabolites of ACs have been found to impair intracellular ironhomeostasis, which provides new perspectives on the role of iron in cardiotoxicity of ACs.149 Figure 8. Schematic structures of daunomycin (R ¼ H) and adriamycin (R ¼ OH).


Figure 9. Crystal structure of nogalamycin2 d(TGTACA)2 complex in which the anthracycline (AC) rings are intercalated into DNAbase pairs.The complex is packed in the crystal lattice as a dimer with the monomers nearly perpendicular to each other.The DNAduplex is shown in ribbon structure in one of the subunits.
ACs are known to bind various metal ions, including transition metal, main group, lanthanides, and uranyl ions.150,151 A number of articles reported that some metal ions, e.g., Fe2þ /3þ, Cu þ /2þ,Pd2þ, Pt2þ, and Tb3þ, play an important role in altering the biochemical properties of ACs.152–155 Thesestudies point a new direction in the pursuit of chemotherapeutic efficacy and lowering toxicity of theseantibiotics. The binding of metal ions may cause a significant influence on the redox property of thesedrugs as shown in their Yb3þ complexes,156 thus affecting their activities. The interactions of DNAand other cell components with metal –AC complexes, and their subsequent damage by the ACcomplexes of redox-active metal ions, including iron and copper,157,158 have been previously studiedby the use of various physical and biochemical methods. Adriamycin has been suggested not toundergo flavo-associated reduction upon intercalation.159 However, a site-specific modification ofDNA bases suggests a possible binding through intercalation,157b although specific electrostaticinteractions cannot be completely ruled out. Pulse radiolysis studies indicate that adriamycinsemiquinone can mediate a long-range electron transfer to as far as 100 base pairs in DNA,160 whichmay also serve as a mechanism toward DNA base modification.
2. Fe–AC Complexes It has been shown that several different metal ions, including alkaline earth metals,161 first-row161 andheavy154 transition metals, and rare earth metals,155,161,162 can bind AC antibiotics very tightly inaqueous and methanol solutions, with the metal bound to one or both of the two b-ketophenolatemoieties depending on the solution conditions. Iron is involved in the actions of several antibiotics,such as BLM discussed in Section 2.A ‘‘Bleomycin'', SN in Section 2.C ‘‘Streptonigrin'', andpossibly ACs,153 which serves as a redox center and can generate free radicals in the presence ofdioxygen under reduction conditions which can damage cell components. The binding of Fe3þ withdaunomycin has been studied by the use of 57Fe Mo¨ssbauer, EPR, and X-ray absorptionspectroscopies, in which several different complexes are seen at mM drug concentrations.153a,163Despite the similar structures of daunomycin and adriamycin, the Mo¨ssbauer spectra of their Fe3þcomplexes are noticeably different which has been attributed to their slight difference in structure andreactivity.164 The binding of Fe3þ with several other ACs has recently been revisited.165 The results suggest that Fe3þ binds these drugs to form 1:1 Fe–drug complexes with the metal bound at 11,12-b-ketophenolate site, and 2:1 Fe2–drug complexes with the metal bound at both b-ketophenolate sites.
The formation of mononuclear, dinuclear, and polynuclear metal –AC complexes are also suggested.
The Fe3þ complexes of these drugs are very complicated systems since their spectra are dependentupon the preparation procedure, equilibrium time, metal-to-drug ratio, and drug concentration.163–165 Different complexes are also formed for lanthanide(III) binding with AC antibiotics observed in anearly study, which is discussed in the next section.
A 1:2 Fe3þ–adriamycin complex was proposed to form a stable complex with calf-thymus DNA in solution. This drug –Fe–DNA tertiary complex is distinct from both the Fe3þ–adriamycin complexand the DNA-intercalated Fe3þ-free adriamycin on the basis of optical and chromatographicstudies.157b In another study, Fe3þ–ACs have been suggested not to intercalate into DNA basepairs until the Fe3þ ion is released, despite the strong binding of Fe3þ with the drugs.157c A recentmutagenesis study indicated that Fe3þ is directly involved in the mutagenicity caused by doxorubicinthrough oxidative DNA damage, which further strengthens the role of Fe in AC action.166 Formationof intracellular Fe–AC complexes have also been confirmed with different methods.167 Furtherstudies are still needed to clarify the mechanistic and structural roles of Fe in the action of this familyof antibiotics.
3. Lanthanide–AC Complexes Lanthanide(III) ions (Ln3þ) have been very widely utilized as substitutes and spectroscopic probes168for biological Ca2þ owing to their very similar ionic radii, binding properties, and coordinationchemistry, yet with much higher affinity constants because of the higher charges of Ln3þ ions (thus isable to probe weak Ca2þ interactions).169 Indeed, both Ln3þ and Ca2þ ions have been reported to bindACs, in which Ln3þ ions show > 3 orders higher in affinity constants.156 Early NMR studies of theparamagnetic Yb3þ–daunomycin complex did not yield useful information for the description of thecoordination chemistry of the complex because of the formation of a mixture and the lack of fullassignment of the paramagnetically shifted 1H-NMR features.170 The binding of several Ln3þ ions,including Pr3þ, Eu3þ, Dy3þ, and Yb3þ, with ACs in both aqueous and methanol solutions underdifferent conditions has recently been revisited by means of electronic spectroscopy, cyclicvoltammetry, and NMR techniques.156,171 Like in the case of Fe3þ-bidning to AC drugs, different complexes are also formed for lanthanide binding with ACs. A 1:1 Yb3þ–daunomycin complex has been successfully prepared in solution,and its hyperfine-shifted 1H-NMR spectrum fully assigned by means of 2D-NMR techniques(Fig. 10).156,171 On the basis of the conclusive signal assignment, the configuration of the complex insolution has been determined to be similar to that of the metal-free drug in solution172 and in thecrystal structure,173 and the metal binding site determined to be the 10,11-b-ketophenolate moiety.
The AC drugs can bind Ln3þ to form complexes in solution with metal-to-drug ratios of 1:1, 1:2, 1:3,and 2:1 depending upon proton activity in the solution.156 All of the complexes have beencharacterized by means of 2D-NMR techniques.156 The complication in the earlier NMR studies hasbeen attributed to the formation of the different complexes at different proton activities, which islikely to be the case for other metal complexes of the ACs.
4. Interactions of ACs and Their Metal Complexes With Other Biomolecules In addition to their DNA intercalation and redox activity, AC antibiotics have been observed tointeract with other biomolecules that may also influence cell functioning and may be the cause of theside effects of these drugs. For example, (a) adriamycin and its Fe3þ, Cu2þ, and Co2þ complexescan cause influence on effector cells of humoral and cell immune response.174 (b) Fe–adriamycincomplex was found to damage erythrocyte ghost membranes, which is attributable to the productionof superoxide and hydrogen peroxide by the complex.157a (c) The Fe2þ, Cu2þ, and Co2þ complexes ofadriamycin are potent inhibitors of propanolol-induced Ca2þ-dependent Kþ efflux, but not Pb2þ-dependent Kþ efflux, whereas Fe3þ–adriamycin can activate Kþ permeability of erythrocytes.
However, the AC rings of adriamycin alone enhances Ca2þ-dependent Kþ efflux from erythrocytes.
These influences are attributed to the influence on cellular Ca2þ transport rather than direct action onKþ channels.175 (d) AC drugs can cause poor wound healing176 as a result of impaired biosynthesis of Figure 10. The 1H-EXSY-NMR spectrum of 1:1 Yb3þ daunomycin complex in methanol. The signals due to the metal complex(top trace) can correlate with those of the free drug (trace onthe left) inthis spectrum, shown as cross peaks inthe‘ 2D map' (labeledwith numbers corresponding to the structure in Figure 8).The spectrum obtained in aqueous solution exhibits similar features as inmethanol.
collagen.177 The inhibition of AC drugs against the Mn2þ-containing prolidase has been observed tobe parallel to the impairment of collagen synthesis.178 The binding of the AC drugs to the Mn2þ in theactive site of prolidase has been suggested to be the cause of the inhibition. Moreover, the higherMn2þ-binding affinity of daunomycin than that of adriamycin has been considered to contribute toits greater potency in inhibition of collagen biosynthesis. (e) The Fe3þ–adriamycin complex isdetermined to be a potent inhibitor of protein kinase C,179 and the Cu2þ–AC complexes are con-sidered to serve as a vehicle to carry Cu2þ to protein kinase C which results in inhibition of theenzyme.180 Direct binding of the complexes with the enzyme has been ruled out. These studiessuggest that the interactions of AC drugs with different bio-targets must be taken into considerationfor further drug design and future studies of the bioactivity and toxicity of these drug family.
E. Aminoglycosides Aminoglycosides form a unique and structurally diverse family of antibiotics (Fig. 11), which includethe famous Waksman's streptomycin and the widely used neomycin (an ingredient in ‘‘tripleantibiotic'' ointment along with bacitracin and polymyxin B). Despite their nephrotoxicity andototoxicity, these antibiotics have remained their clinical values and also serve as lead drugs forrational design of next-generation antibiotics.181 1. RNA-Binding and Aminoglycoside Action Aminoglycoside antibiotics are known to bind RNA which is considered the key mechanism in theirantibiotic activities.181–183 This binding decreases translational accuracy and interferes withtranslocation of the ribosome.184 For example, neomycin-like aminoglycosides bind rRNA near theaminoacyl site, preventing chemical modification on the nucleotides in the aminoacyl site.185Neomycin B has been determined to bind to the transactivation-responsive element of HIV-1 RNA.186Neomycin has also been determined to inhibit the self-cleavage of the ribozyme from human hepatitisd virus by direct replacement of the active divalent metal ions.187 Moreover, aminoglycosides areknown to bind and cleave hairpin ribozyme in the absence of Mg2þ, however, with much smaller rate Figure 11. Schematic structures of aminoglycosides (A) neomycin B (R ¼ NH2) and paromomycin (R ¼ OH), (B) gentamicin C1(R1 ¼ R2 ¼ CH3), gentamicin C2 (R1 ¼ CH3; R2 ¼ H), and gentamicin C1A (R1 ¼ R2 ¼ H), and (C) kanamycin A (R1 ¼ OH;R2 ¼ OH), kanamycin B (R1 ¼ OH; R2 ¼ NH2), and tobramycin (R1 ¼ H; R2 ¼ OH).
constants kcat in most cases except neomycin and apramycin of 18 and 13 times smaller, respectively,than that of Mg2þ-catalyzed cleavage.188 The interaction of aminoglycosides with RNA has been investigated by the use of small RNA nucleotides that contain the drug recognition site.189 The solution structure of a 27-mer RNAmolecule, and the structures of this RNA nucleotide bound with paromomycin and gentamicin havealso been determined with NMR spectroscopy.190 The structures of an E. coli decoding region A-siteoligonucleotide with and without a bound paromomycin have also been resolved by means ofhomonuclear and heteronuclear NMR techniques, wherein the two structures are found similar exceptat the antibiotic binding region.190,191 Crystal structures of ribosomal 30S RNA subunit192 and itscomplexes with paromomycin, streptomycin, and spectinomycin have been resolved (Fig. 12, Top).193These structures have provided structural details about the conserved A1492 and A1493 region aswell as detailed interactions of aminoglycoside antibiotics with RNA (Fig. 12, Bottom) which affordstructural basis for the understanding of the action of aminoglycosides. Another crystal structure ofRNA –aminoglycoside complex has also been recently determined, wherein two ribosomal decodingA-sites are bound with two paromomycin molecules.194 In both solution and crystal structures, therings A and B of the antibiotics (cf. Fig. 11) are found to be involved in specific interactions with RNAvia H-bonding with G and A nucleotides, whereas rings C and D in paromomycin and neomycincontribute to the drug binding affinity to RNA. Consequently, methylation of G or A nucleotide canlead to different bacterial resistances to this family of drugs.195 Nevertheless, different drugs are foundto bind at different locations in 30S RNA, which could be metal-dependent (cf. Section 3 for TCbinding to RNA).
2. Metal Binding and Bioactivities Metal ions have been determined to be involved in some unique activities of aminoglycosides. Thebinding of iron to gentamicin (Fig. 11) has been postulated to induce free radical formation which


Figure 12. Crystal structure of antibiotic-bound 30S rRNA complex (top structure, Protein Data Bank ID 1FJG.pdb) and the detailsabout the paromomycin-binding environment (bottom).The RNA molecule is shown in green, proteins in gray, and the aminoglyco-side antibiotics paromomycin, streptomycin, and spectinomycin are shown in purple, blue, and red colors, respectively in the topstructure.There are 96 Mg2þ ions and 2 Zn2þ ions (bound to the peptide chain) found in this structure (not shown in the figure); how-ever, the metal ions are not involved in the binding of the antibiotics.
causes peroxidation of lipids.196 The Fe2þ/3þ complexes of gentamicin have recently investigatedwith NMR, in which a low-spin 2:1 drug-to-Fe2þ complex as well as a 1:1 and a 2:1 drug-to-Fe3þcomplexes have been proposed to form.197 These redox-active iron complexes were implied foraminoglycoside toxicity.
The macrolide antibiotic erythromycin has a structure different from the streptomycin-like antibiotics, yet it contains two sugar moieties (one being a t-aminosugar), carbonyl, and hydroxylgroups which potentially can serve as metal binding ligands. An erythromycin–iron complex wasobserved to exhibit superoxide scavenging activity that was not seen for the antibiotic without themetal.198 However, the physical and structural properties of the metal binding site and the structure ofthe complex were not determined in the study.
Several other aminoglycoside antibiotics have been determined to bind Cu2þ, including lincomycin,199 kasugamycin,200 kanamycin B,201 tobramycin,202 genticin,203 and the semi-syntheticamikacin204 (Fig. 11). In addition, a few simple amino sugars have also been reported to bind Cu2þ,which serve as simple model systems for metal-binding of aminoglycoside antibiotics.205 In all thecases, the binding of Cu2þ to the aminoglycosides are highly pH-dependent, and afford multi-speciesaround neutral pH based on the results from potentiometric and EPR studies. The Cu2þ–amino-glycoside complexes are observed to exhibit oxidative activity, which can catalyze oxidationof nucleotides in the presence of H2O2.199–201,204 Hydrolytic cleavage of DNA206 and RNAmolecules207 and the RNA of the HIV-1 viral Rev response element207 under physiological conditionsby Cu2þ–aminoglycoside complexes was also observed. The metal ion in these complexes has been proposed to bind to the drugs through a chelating vicinal aminohydroxyl binding moiety of the drugs.
The binding site of Cu2þ in kanamycin A has been determined to be the 3 0 –NH2 and 40–OH groups ofring C (Fig. 11C) by means of 13C-NMR relaxation and potentiometric measurements.201,206 Quinolones are comprised of a large family of antibacterial agents such as nalidixic acid, pefloxacin,norfloxacin, ofloxacin, and ciprofloxacin (Fig. 13).208,209 The first-generation nalidixic acid is activeonly against Gram-negative bacteria, whereas the later generations, such as the fluoroquinolones witha fluorine atom on the number 6 carbon (Fig. 13B), have been modified to become effective anti-bacterial agents which exhibit a broad spectrum of activity highly against Gram-negative bacteria andless active against Gram-positive bacteria and also show significant activity against anaerobicbacteria. Fluoroquinolones have been further modified to produce quinobenzoxazines (Fig. 13C),which are found to show anti-tumor activities (whereas the parent quinolones lack such activities)believed to be attributable to their interaction with topoisomerase II.210 Ciprofloxacin (Cipro1 ofBristol-Myers) is a prototypical fluoroquinolone which has been brought on the stage in recent anti-bioterrorism reactions. It has become ‘‘the antibiotic of choice'' for fighting against anthrax caused byBacillus anthracis prior to the release of significant amount of toxin by the bacterium, despite the factthat several other antibiotics are also effect against this bacterium.211 Since this family of drugs havebecome widely used, one should also bear in mind the risk of serious side effects such as tendinopathyas a consequence of quinolone treatment.212 1. Metal Complexes of Quinolones Quinolones can bind several divalent metal ions, including Mg2þ, Ca2þ, Mn2þ, Fe2þ /3þ, Co2þ, Ni2þ,Cu2þ, Zn2þ, Cd2þ, and Al3þ,213,214 and may result in change in their activity. Mg2þ and Al3þ werefound to decrease the activity of the drugs,215 whereas Fe3þ and Zn2þ complexes were found toexhibit higher activities.216 The crystal structures of the Ni2þ and Cu2þ complexes of cinoxacin andciprofloxacin have been solved, in which the metals are found to bind to the a-carboxylketo moiety toform 1:2 metal-to-drug complexes.214 The complexes have a pseudo-axial symmetry with the twodrug ligands bound symmetrically at the equatorial positions. The axial symmetry is also seen in theEPR spectra of Cu2þ–(drug)2 complexes.213e,214c The drug was also determined by means ofcrystallography to form a 1:3 Co2þ :drug3 complex.217 A few metal complexes (Fe3þ, Cu2þ, and Bi3þ)of quinolones were prepared in acidic solutions, from which crystals were obtained and structuressolved.218 However, the metal ions in these crystals do not bind directly to the drugs owing toprotonation of the carboxylate group, which may not be relevant to the drug action under Figure 13. The structures of (A) nalidixic acid; (B) the prototypical fluoroquinolones (F substitution at position 6) ciprofloxacin (Cipro ); R1 ¼ H; R2 ¼ cyclopropyl, norflozacin; R1 ¼ H; R2 ¼ ethyl, and pefloxacin; R1 ¼ CH3; R2 ¼ ethyl; and (C) a prototypicalquinobenzoxazine, A-62176.
physiological conditions. The formation of M2þ (quinolone)(2,2 0-dipyridine) ternary complexes(M ¼ Co, Ni, and Cu) was observed by means of electrospray ionization and laser desorption massspectroscopy.219 A recent theoretical study suggested that metal binding to these drugs is associatedwith the action of these drugs, and fluorescence quenching measurements indicate the presence of ap–p stacking which has been suggested to be associated with the DNA intercalation capacities of thedrugs and their Cu2þ complexes.220 2. Mechanism of Quinolone Action The binding of quinolones to DNA-gyrase or topoisomerase IV has been considered the key step in theaction of these drugs, which prohibits DNA religation activity and distorts DNA in the complex.221Recent studies on the mapping of the functional interaction domain of topoisomerase II revealed thatthe quinolone-action site on the enzyme overlaps with those sites for the DNA cleavage-enhancingdrugs, including etoposide, amsacrine, and genistein.222 DNA has been considered the target forquinolone drugs, and a cooperative quinolone–DNA binding model of DNA gyrase in the presence ofATP is proposed.223 Norfloxacin exhibits a Mg2þ-dependent binding to plasmid DNA in the absenceof the enzymes,224 wherein metal –drug, metal –DNA, and drug –metal –DNA complexes aredetected. The drug does not bind to DNA in the absence or in the presence of an access amount ofMg2þ. Intercalation of norfloxacin into DNA is proposed in the study, and Mg2þ is proposed to serveas a ‘‘bridge'' for the carboxylate of the drug to interact with DNA. However, DNA unwindingefficiency of 10 by this drug is only marginal for a weak intercalation.224,225 Fluorine-19 NMRstudy of the binding of pefloxacin with double stranded DNA also revealed the participation of Mg2þin the binding.226 Moreover, the Mg2þ-dependent single-stranded DNA binding affinities of several 6-substituted quinolones are found to correlate with the gyrase poisoning activity of these drugs,227confirming the involvement of Mg2þ in such interaction and the significance of the substitution atposition-6 and supporting the mechanism derived from quinolone–DNA interaction.
Quinobenzoxazines have been proposed to bind DNA duplex in the presence of Mg2þ to form a ternary complex in the form of drug2–Mg2þ –DNA, in which one drug molecule is proposed to intercalate into the DNA base pairs while the other is ‘‘externally bound.''228 The two Mg2þ ions serveas salt bridges which interact with both molecules of the drug and the phosphoester backbone of DNA.
These drugs have been determined to form a 1:1 or 2:2 complex with Mn2þ in methanol by means ofJob plot229 (in which absorption is measured against different metal-to-ligand ratios), which alsoimplies a possible formation of ternary complexes between 2:2 metal –quinolone complexes andDNA. The metal –quinobenzoxazine complex interacts with DNA in a cooperative manner, i.e., a 4:4metal –drug complex is proposed to interact with DNA as a unit, in which two drug moleculesintercalate into DNA base pairs while the two ‘‘external'' drugs have p–p interaction and are expectedto interact with the enzyme topoisomerase II or gyrase. The 2:2 metal –drug complex is also suggestedto be assembled in the presence of topoisomerase II based on the results from photocleavage assay, theuse of mismatch sequences, and competition experiments.230 The formation of the 2:2 metal –drugcomplexes suggests that different quinobenzoxazine or quinolone drug molecules should be utilizedto form ‘‘hybrids'' for the pursuit of optimal structure–activity relationship.
The antibiotic activities of the platinum complexes cis-diamminedichloroplatinum (also commonlyknown as cisplatin, cis-[PtII(NH3)2Cl2]; R1 ¼ NH3 and R2 ¼ Cl in Fig. 14) and cis-PtIVCl6(NH4)2were found serendipitously by Barnett Rosenberg to cause dramatic elongation of E. coli during astudy of the influence of electric fields on the growth of the bacterium by the use of a platinum electrodein a buffer solution containing NH4Cl.231 The abnormal growth of this bacterium was later identifiedto be caused by the oxidation of the Pt electrode to form the Pt(IV) salt, which was confirmed viachemical synthesis of the compound. In the meantime, the Pt(II) compound cis-[Pt(NH3)2Cl2] was Figure 14. Schematic structures of (A) cisplatin (R1 ¼ NH3 and R2 ¼ Cl) and ‘ transplatin' (R1 ¼ Cl and R2 ¼ NH3) and (B) theless toxic analogue carboplatin.
also identified to be a potent antibiotic agent, causing the same effect as the Pt(IV) salt. Thus, thesesynthetic metal complexes can be considered metalloantibiotics from a broad sense of the term as theycan inhibit the growth of microorganisms. After its discovery, cisplatin was soon found to be a potentanti-cancer agent and is nowadays one of the most prescribed anti-cancer drugs which has been usedfor the treatment of several different cancers and tumors, including head and neck tumor andtesticular, lung, breast, and ovarian cancers.232 DNA is considered the main biological target ofcisplatin. The coordination chemistry and reactivity of cisplatin and the interaction of cisplatin withDNA have been extensively studied by means of 1H-, 31P-, and 195Pt-NMR spectroscopy and X-raycrystallography, and has previously been reviewed in a number of publications.233 1. Cisplatin–DNA Complexes The chemistry of cis-[Pt(NH3)2Cl2] has been thoroughly investigated in the late 1890s by AlfredWerner.234 The two bound chloride ions in cisplatin are relatively labile, which can undergo exchangewith nucleophiles such as amine bases. Upon introduction of DNA, cisplatin binds to the N7 nitrogenof two adjacent guanidine bases or guanidine–adenine bases in the major groove, or two proximalguanidine bases on different strands in the minor groove which distorts the DNA structures by bendingthe helix by 40–60 and a helical twist of 25–32. This binding pattern and structural perturbation onDNA have recently been revealed by means of crystallography235 and NMR spectroscopy236 (Fig. 15).
The DNA binding mode of cisplatin cannot be achieved by its stereoisomer ‘‘transplatin'' (R1 ¼ Cl; R2 ¼ NH3 in Fig. 14),237 in which the two trans chloride cannot bind to adjacent guanidinebases as in the case of cisplatin. The lability of the Pt–Cl bond still allows nucleophilic substitution tooccur in transplatin which can result in DNA binding. However, the DNA binding of transplatin issignificantly different from that of cisplatin, wherein cross-strand linkage becomes predomi-nant.233,238 The observation of significant cytotoxicity of the trans analogues with pyridine in place ofthe ammonia239 and high anti-tumor activity of trans-imino analogues240 suggest that ‘‘transplatin''analogues are worth further exploration for design of new platinum antineoplastic agents.237 2. Cisplatin Conjugates Cisplatin has been linked to bioactive molecules to form conjugates which exhibit unique proper-ties in terms of DNA binding and anti-tumor activity. For example, adriamycin (Section 2.D‘‘Anthracyclines'') can form complexes with PtCl2 to afford cisplatin-like complexes,154a such as cis-dichloro-t-butylamine-adriamycino-platinum which has been determined to be active againstmurine carcinoma and leukemia.154b This complex has been suggested to interact with DNA byintercalation of the AC rings rather than covalent binding to the Pt center. A hormone-anchoredcisplatin complex has been prepared in which testosterone is bound to cisplatin in place of thediammine groups via a thiosemicarbazone linkage.241 This conjugate exhibits a higher activity thancisplatin against human breast cancer cell line MCF-7. The binding of cisplatin with proteins,including serum albumin242 and transferrin,243 has also been reported which is considered to playimportant role in the metabolism and bioactivity of this drug. The interaction of proteins with cisplatinmay possibly mediate cell response to the drug, which has been recently reviewed.244


Figure 15. Stereo view of the structures of d(CCTGGTCC)*d(GGACCAGG) obtained with NMR spectroscopy (top structure; ProteinData Bank ID1AU5.pdb), in which cisplatin is bound at the center GG nucleotides of one strand, and d(CCTCGCTCTC)*d(GGAGC-GAGAG) obtained with X-ray crystallography (bottom structure; Protein Data Bank ID1A2E.pdb), in which cisplastin is bound to thecenter G from each strand. The binding of cisplatin to DNA significantly distorts DNA structure, particularly in the case of the cross-strand binding.
3. Cisplatin Analogues A large number of cisplatin-like compounds have been synthesized, their molecular propertiesthoroughly characterized, and their anti-tumor activities evaluated.233,245 Of these new analogues, thecompound carboplatin (cis-diammine-cyclobutane-1,1-dicarboxylatoplatinum, Fig. 14B) exhibitslower toxicity than cisplatin and has currently been used clinically for cancer treatment. The othercompounds such as nedaplatin (cis-diammine-1-hydroxoacetatoplatinum) and oxaliplatin (1,2-diaminocyclohexane-oxalatoplatinum) also exhibit potential antineoplastic activities, which havegained approval for clinical use in some countries and are under extensive evaluation.233,245,246 Several cisplatin analogues with two Pt centers have recently been prepared, possessing a general formula of [(trans-PtCl(NH3)2)2-m-L]2þ in which L is a diamine linker.247 Because of the presence oftwo DNA-binding motifs in each molecule, binding of these dinuclear platinum complexes to DNAduplex affords intrastrand and/or interstrand cross-link, wherein the bending of DNA at the bindingsite is much less than that caused by cisplatin.247,248 These dinuclear platinum compounds exhibitanti-tumor activities differently from cisplatin and may also be different from each other, and are potential new anti-tumor agents. Analogous compounds with multi-platinum centers have also beenprepared and show significant anti-tumor activities and a cellular response different from cisplatin,and have been under clinical trial.249 The DNA-binding pattern of these new compounds has also beeninvestigated which shows a similar bifunctional manner as dinuclear platinum compounds.250 Platinum(IV) complexes have been known to exhibit anticancer activities.231 Several Pt(IV) complexes have entered clinical trials;251 however, they have not been widely used because of loweractivities than cisplatin or high toxicity and viability of drug uptake, including cis,trans,cis-[PtCl2(OH)2(isopropylamine)2] (iproplatin, CHIP, or JM9),252 [PtCl4(D,L-cyclohexane-1,2-diamine)](tetraplatin or ormaplatin),253 and cis,trans-[PtCl2(OAc)2(NH3)(NH2C6H5)] (JM216 or satrapla-tin254). The bioinorganic chemistry of Pt(IV) complexes has recently been extensively reviewed.255 H. Organometallics Organometallic compounds are a large family of unique synthetic metal-containing organiccompounds, which are characterized by the presence of direct metal –carbon bond(s). Severalorganometallic compounds have been found to exhibit antineoplastic activities.256 Of these, the‘‘metallocene'' compounds M(IV)Cp2Cl2 (Cp ¼ cyclopentadienyl; M ¼ Ti, V, Nb, and Mo) showsignificant activities toward several experimental animal tumors and human tumors on nude mice,whereas the Zr and Hf analogues do not show anti-tumor activity.257 The Ti compound has enteredclinical trials.258 In addition to the metallocenes, there are a number of non-platinum metal complexeswhich have been extensively studied and tested for their anti-tumor activities and are covered in recentreviews.233e,256 Although metallocenes were originally considered to bind DNA similar to cisplatin, recent studies indicated that they do not bind tightly to DNA at neutral pH.259,260 Nevertheless, DNA is still abinding target of these compounds under certain conditions,256 as suggested by NMR studies.261TiCp2Cl2 has been suggested to exhibit an anti-tumor mechanism different from cisplatin,262 showinginhibitory activity toward protein kinase C and DNA topoisomerase II.259 The hydrolysis of thesecompounds into M(IV)Cp has been proposed to render their anti-tumor activities.259 The a values of the bound water molecules in MCp2(H2O)2 result in different charges on the compounds, which relates to their capability of entering cells. The high acidity of TiCp with pKa values of 3.51 and 4.35, which afford a neutral species at pH 7.0, and its reasonable stabilitywith t1/2 ¼ 57 hr for Cp dissociation263 may account for its high anti-tumor activity.
MCp2Cl2 can form conjugates with adriamycin (Section 2.D ‘‘Anthracyclines'') to give 1:2 metal-to-drug complexes (M ¼ V and Zr) and 1:1 and 1:2 complexes (M ¼ Ti).264 While the Zrconjugate does not show activity toward P-388 leukemia, the Ti complexes exhibit activitycomparable to the free drugs. The structures of these metal conjugates were not determined in theprevious study. The metal ions are suggested to assist the binding of the drug to DNA and red bloodcell membrane. However, these metal complexes do not catalyze electron transfer from NADH todioxygen as does adriamycin (Section 2.D ‘‘Anthracyclines''), which possibly may decrease thecardiotoxicity of adriamycin. Thus, these conjugates seem to serve as bifunctional anti-tumorcompounds, i.e., to release adriamycin and the M–Cp2 complex.
The tetracyclines (TCs) have once been widely used as both external and internal medicines for anextended period of time because of their broad-spectrum activity toward both Gram-positive and-negative bacteria, and also their activity toward rickettsiae, chlamydiae, and protozoans, such as theprototypical TC aureomycin (Fig. 16) produced by Streptomyces aureofaciens.265 The antibioticactivity of TCs is attributed to their binding to the ribosome which inhibits protein synthesis.266 Theirusage has been limited in recent years because of side effects, including staining of teeth and increase Figure 16. Schematic structure of aureomycin (7-chlortetracycline). Substitute OH for 5-H and H for 7-Cl afford terramycin (5-oxytetracycline).
in bacterial resistance. However, recent studies of the mechanism for bacterial resistance of this drughas afforded new insight into rational design of analogues and searching for new analogues of thisbroad-spectrum antibiotic family, such as the novel 9-glycylamido derivatives the ‘‘glycylcyclines,''for defending bacterial infections.267,268 One of the glycylcyclines 9-t-butylglycylamido-minocy-cline (GAR-936, tigilcycline) is currently under phase II clinical trials.269 The metal-binding capability of TCs has been well documented,150 including the binding withalkaline earth and transition metal ions (VO2þ, Cr3þ, Mn2þ, Fe2þ/3þ, Co2þ, Ni2þ, Cu2þ, and Zn2þ) andAl3þ .270–274 TCs have been determined to be present mainly as Ca2þ -bound form (and Mg2þ -boundform to a lesser extent) in the plasma when they are not bound to proteins such as serum albumin.
Thus, the bio-availability of TCs should be dependent upon the physical and biochemical properties oftheir metal complexes instead of their metal-free forms. Metal binding to different TCs are found to beslightly different which has been suggested to be correlated to their pharmacodynamic effect.271b The acidic oxy-groups at positions 1, 3, 10, 11, and 12 of TC are the potential metal binding/ chelating sites (Fig. 16). The acidity of these groups has been determined to follow the order of 3–OH > 12–OH > 4-ammonium > 10–OH.150 The 11,12-b-ketoenol moiety has been considered tobe the primary metal binding site,275 which has also been determined to be the Mg2þ binding site inthe repressor TetR –Mg2þ–TC ternary complex (see later). A recent study indicated that TC forms 2:1TC:metal complexes with 3d transition metal ions in non-aqueous solutions, in which the metal isbound at the 2-amido 3-enol chelating site.270 Moreover, the formation of metal –TC complexes withdifferent stoichiometries, including 2:1, 1:1, and 1:2 metal:TC ratios, has also been suggested in theprevious studies.
The antibiotic activity of TCs is attributed to their binding to the ribosome.266,276 TCs have beenreported to bind different forms of RNA, including the ribosome, bulk RNA, rRNA, and ribozymes.
The studies of TC binding and interaction with RNAs have recently been reviewed.277 The binding ofTCs to bulk RNA is not specific, and may not be significant for their antibiotic activity.278 On the otherhand, this family of antibiotics bind to the ribosome at the 30S subunit with Kd of 1–20 mM(in addition to many other low affinity sites). This binding induces a conformational change thatprevents tRNA from binding to the ribosome and results in interference of protein synthesis.279 Theinteraction of TCs with 16S rRNA has recently been extensively studied with photo-modification,activity assay, mutation, and other methods, from which the TC-binding sites have beenidentified.277,280 The crystal structures of TC-bound small ribosomal subunit have recently been resolved,281 further confirming the significance of such binding in the action of this antibiotic family. Two TCmolecules are found to bind to the RNA (Fig. 17).281a One of the TC molecules may involve a Mg2þion (at the 11,12-b-ketophenolate site that is found in metal binding studies of TC discussed above) in



Figure 17. Top: The crystal structure of 30S rRNAwith two tetracycline (TC) molecules bound (red).There are 96 Mg2þ ions found inthe structure. The one located near 11,12-b-ketoenol moiety may be involved in the binding of the drug to RNA (bottom enlargedbinding site of theTC on the right), as in the case of the binding of the drug toTetR receptor discussed below.
Figure 18. Top: Stereo view of one subunit of the ternary complex formed between class DTetR repressor and Mg2þ aureomycin(Protein Data Band ID 2TCT.pdb). The Mg2þ aureomycin complex is shown in red. The DNA-binding domain is located at the N-terminus on the left.The 2-fold symmetry of theTC TetR dimeric complex allows the binding of the complex to the15-base pair tet-operator. Bottom: Structure of theTC biding site in theTetR Mg2þ TC ternary complex. The drug complex is bound to the proteinthrough His100 via Mg2þ (green), and is also H-bonded with the protein through three amino acid side chains.
binding to RNA through the phosphates of C1054, G1197, and G1198 (Fig. 17). The TC moleculesoccupy the sites that are distinct from those for the binding of aminoglycosides discussed in Section2.E ‘‘Aminoglycosides'' (cf. Fig. 12).
Inhibition of ribozymes by TCs has been studied, including groups I and II introns, hammerhead ribozyme, a ribozyme from hepatitis delta virus, and Neurospora crassa Varkud satellite RNA.187,282The concentration for 50% inhibition (IC50%) of ribozymes has been determined to be 10–500 mMfor several different TCs, with hydrophobic TCs showing higher inhibitions. The large IC50% valuesindicate that these drugs are weak inhibitors for ribozymes, or may even serve as non-specificinhibitors.277 The binding sites, the binding nature, the pattern for the inhibition, and the role of metalions (particularly the RNA-significant Mg2þ ) in the binding with ribozymes were not revealed in theprevious studies.
C. Metal-Dependent Bacterial Resistance Despite the high potency as broad-spectrum antibiotics, TCs are of little use nowadays because oftheir bacterial resistance.265b,283 The predominant TC-resistance mechanism in Gram-negativebacteria is active efflux of the drugs mediated by the antiporter membrane protein TetA which pumpsout TC as a Mg2þ complex coupled with proton uptake.284 The expression of TetA is controlled by therepressor protein TetR, whose binding to operator prevents transcription of both tetR and tetA genes.
A conformational change of the TetR repressor is supposed to occur upon binding of TC in thepresence of divalent metal ions.285 The conformational change results in the release of the repressorfrom the operator and initiates the expression of TetA for active TC efflux. The crystal structures of therepressor TetR and the ternary complex TetR–Mg2þ–TC have been resolved which confirm theinduction of the conformational change of the repressor upon the binding of the Mg2þ–TCcomplex.286,287 A structure of Mg2þ–TC-bound TetR is shown in Figure 18(Top). TetR is a dimeric protein with 10 a-helical structures, of which the first three helical bundles from the N-terminus of each subunitserve as the DNA binding site. The tet-operator is composed of 15 base pairs shown below, which has a2-fold symmetry (boxed sequences) with respect to the center T-A base pair.
Upon Mg2þ -TC binding, significant conformational changes of TetR are observed, including changes in the drug binding site and the DNA binding site.286 Significant changes are also observedat helix-9, suggesting that the opening at the C-terminus of helix-9 serves as the entrance for thedrug as this opening is significantly narrowed after TC binding.286,287 Mg2þ in the ternary TetR–Mg2þ–TC complex is found to bind to the drug at the 11,12-b-ketoenol moiety (as suggested in earlymetal-binding studies, see above) and to TetR via His100, in addition to three water molecules(Fig. 18).
Fe2þ can form a ternary complex with TC and TetR in place of Mg2þ .285,288 An in vitro induction assay shows that Fe2þ–TC is a stronger inducer of Tet repressor than Mg2þ–TC by more than 1,000times, suggesting that Fe2þ may play a role in TC resistance in vivo.288b Specific sites of cleavage ofTetR by the bound Fe2þ is achieved in these studies via Fenton chemistry, and have been identified bymeans of electrospray ionization mass spectroscopy. The cleavages are found to occur at residues104 and 105, 56 and 136, and 144 and 147 in order of preference. This cleavage pattern is consistentwith the geometric locations of the respective residues to the metal center found in the crystalstructures. The determination of the roles of metal ion in the binding of TC to TetR and in the structureof the TC–M2þ–TetR complex is expected to lead to rational design of TC analogues that exhibit broad-spectrum antibiotic activities yet devoid of bacterial resistance, such as the glycylcyclinefamily.267–269 Bacitracin is a metal-dependent dodecapeptide antibiotic excreted by Bacillus species, including B.
subtilis and B. licheniformis. It is a narrow spectrum antibiotic directed primarily against Gram-positive bacteria, such as Staphylococcus and Streptococcus, via inhibition of cell wall synthesis.289Currently, this antibiotic is commercially produced in large quantities as an animal feed additive forlivestock290 and in human medicinal ‘‘triple antibiotics'' ointments (along with polymyxin andneomycin).291 The historical perspectives, the structure of metallobacitracin, and the structure–function relationship of this antibiotic have recently been reviewed.292 A. Congeners and Biosynthesis This antibiotic is produced as a mixture of many closely related peptides, in which bacitracins A1 andB1 are the major components with the most potent activity (Fig. 19).293 Several congeners of thisantibiotic have previously been isolated and characterized by means of amino acid sequence andmass spectroscopy.293 Bacitracin contains four D-amino acids, including D-Glu4, D-Orn7, D-Phe9,and D-Asp11, and a thiazoline ring formed by condensation of the carboxylate of Ile1 with the –NH2and the –SH groups of Cys2 (Fig. 19). A cyclic heptapeptide structure is formed via an amide bondlinkage between the side-chain e-NH2 of Lys6 and the C-terminus of Asn12. These unusual structuralfeatures may protect this peptide from degradation by proteases.
Like those structurally diverse peptides and polyketides and their hybrids such as BLM (Section 2.A ‘‘Bleomycin''), bacitracin congeners are also nonribosomal products of a large peptide synthetasecomplex.294 The structure and mechanism of bacitracin synthetase resemble those of other peptideand polyketide synthetases, which are comprised of a multi-domain modular structure for thecatalysis of the initiation of the thioester linkage to the enzyme, elongation of the thioester-linkedamino acid, and termination of the peptide or polyketide chain by a thioesterase domain.77 Bacitracin synthetase has been known since its early studies to comprise of a complex modular structure as in the case of other peptide/polyketide synthetases. This enzyme catalyzes an ATP-dependent synthesis of bacitracin, starting from the N-terminus based on the observation of a few N-terminal peptidyl intermediates such as Ile-Cys, Ile-Cys-Leu, Ile-Cys-Leu-Glu, and several otherN-Ile-containing peptides.295 The role of ATP has been suggested to be involved in the formation ofthe labile aminoacyl adenosine intermediates. As in the case of other nonribosomal peptide/polyketide biosyntheses, the synthesis of bacitracin has been suggested to involve thioester-linkages Figure 19. Amino acid sequences and N-terminal structures ofa few congeners of bacitracin. Bacitracin A1contains Ile5,Ile8, and R1;A2, Ile5, Ile8, and R2; B1, Ile5, Ile8, and R3; B2,Val5, Ile8, and R1; B3, Ile5,Val8, and R1; F, Ile5, Ile8, and R4; E,Val5,Val8, and R3; H1, Ile5, Ile8, andR5; H2,Val5, Ile8, and R4; H3, Ile5,Val8, and R4; I1,Val5, Ile8, and R5; I2, Ile5,Val8, R5; I3,Val5,Val8, and R4.
based on the observation of covalent peptide –enzyme complexes.77 The thiazoline ring in bacitracinhas been suggested to be synthesized at the stage of Ile-Cys formation on the basis of the detection ofthe oxidized thiazole product of the Ile-Cys intermediate.296 The thiazoline ring and analogousthiazole ring are found in a number of peptide antibiotics and siderophores that are synthesized with asimilar mechanism.76,297 An early study revealed that the activity of bacitracin synthetase is affectedby Mg2þ , Mn2þ , Fe2þ , and Co2þ (Zn2þ was not checked) as well as bacitracin,298 suggesting afeedback control of the synthetase by bacitracin and metal ions.
The bacitracin synthetase operon contains the gene bacA, bacB, and bacC which have been recently cloned and determined to encode three products BA1 of 598 kDa, BA2 of 297 kDa, and BA3of 723 kDa.294 BA1 contains five modules to incorporate the first five amino acids, an epimerizationdomain attached to the forth module for the inclusion of D-Asp4, and a cyclization domain for theformation of the thiazoline ring between Ile1 and Cys2; BA2 is comprised of two modules and anepimerization domain for D-Orn6 incorporation; and BA2 contains five modules for the addition ofIle8-Asn12 with two epimerization domains and the thioesterase domain, consistent with previousstudies.294 A disruption of the bacB gene results in a bacitracin-deficient mutant, confirming theinvolvement of this gene in bacitracin synthesis. Moreover, the expression of a foreign bacitracinsynthetase in a host B. subtilis results in the production of bacitracin at high level, confirming thefunctional role of bacitracin synthetase and its association with bacitracin self-resistance genes.299The available genes of bacitracin synthetase and other peptide/polyketide synthases/synthetasesafford us the tools for possible preparation of different congeners of peptide antibiotics with higheractivities for combating bacterial infections.79 B. Metal Complexes and Antibiotic Mechanism Bacitracin requires a divalent metal ion such as Zn2þ for its antibiotic activity,300 and can form a 1:1complex with several divalent transition metal ions, including Co2þ , Ni2þ , Cu2þ , and Zn2þ .301 TheCo2þ–bacitracin complex binds tightly to C55-isoprenyl (undecaisoprenyl or bactoprenyl) pyro-phosphate with a formation constant of 1.05  106 M1.302 This binding capability of metall-obacitracin presumably prevents the long-chain pyrophosphate from dephosphorylation by amembrane-bound pyrophosphatase, which subsequently inhibits cell wall synthesis because thehydrolytic product undecaisoprenyl phosphate is required to covalently bind UDP-sugars fortransport of the sugars during cell wall synthesis.303 Thus, the binding of metal –bacitracin complexesto undecaisoprenyl pyrophosphate is the key step in the inhibition of cell wall synthesis by thisantibiotic since the sugars become unavailable as building blocks during cell wall synthesis. Althoughthe formation of the Co2þ–bacitracin–undecaisoprenyl pyrophosphate ternary complex wassuggested in previous studies,302 the structure of different metal –bacitracin complexes and thestructure –activity relationship of this antibiotic were not conclusively defined.
C. Coordination Chemistry of Metal Complexes An early NMR study of Zn2þ–bacitracin suggested that His-10 and the thiazoline ring sulfur atomrather than the nitrogen atom were coordinated to the metal,301 but did not implicate other moieties asmetal binding ligands. A later EPR study on Cu2þ–bacitracin revealed a slightly rhombic EPRspectrum with gx ¼ 2.058, gy ¼ 2.047, and gz ¼ 2.261 and a large copper hyperfine couplingconstants of Az ¼ 534 MHz, typical of a tetragonally distorted Cu2þ center.301b The detection of clearsuperhyperfine coupling is indicative of the presence of N ligands. The coordination environment ofthis complex was suggested to be comprised of ligands from thiazoline ring nitrogen, the His10imidazole, the D-Glu4 carboxylate, and the Asp11 carboxylate. The results from a recent X-rayabsorption spectroscopic study via the examination of the extended X-ray absorption fine structure(EXAFS) of Zn2þ–bacitracin in solid form indicated the involvement of three nitrogens and oneoxygen in the first coordination sphere with a tetrahedral-like geometry.304 The ligands are suggested to be thiazoline nitrogen, His10 imidazole, D-Glu4, and possibly the N-terminal amino group. Thebinding of metal through thiazoline sulfur is excluded in this study. This study has provided furtherinsight into the metal binding environment of Zn2þ–bacitracin and corroborates some previousobservations.
The structure of the metal coordination has emerged from the spectroscopic studies discussed above and a recent NMR study of the paramagnetic Co2þ complex.305 The hyperfine-shifted 1H-NMR signals of Co2þ–bacitracin complex have been successfully assigned by the use of both 1D- and2D-NMR techniques as shown in Figure 20. The metal-binding ligands have been conclusivelyidentified in this NMR study, which are assigned to be the Ne of His10, the carboxylate of D-Glu4, andthe thiazoline nitrogen. The N-terminal amino group is not bound to the metal. The identification ofseveral signals attributed to protons near the metal allows a model to be built using relaxation times asdistance constrains (Fig. 21). It is interesting to note in the model that a hydrophobic pocket is formedby the side chains of Ile5 and D-Phe9, which presumably can serve as the binding site for thehydrophobic hydrocarbon chain of the sugar-carrying undecaisoprenyl pyrophosphate.
D. Structure–Function Relationship Further investigation of the Co2þ complexes of several other congeners, including the activebacitracins B1 and B2 and the inactive stereoisomer A2 and the oxidized form F (which have beencharacterized with mass spectrometry and 1D- and 2D-NMR techniques306), revealed that a propermetal binding is essential for bacitracin to exhibit antibiotic activity. That is, all the active congenershave similar metal binding properties and coordination chemistry as bacitracin A1, whereas the metalbinding patterns of the inactive bacitracins A2 and F are different from that of the active forms, inwhich Glu10 in bacitracin A2 and both Glu10 and the oxidized thiazole ring in F are not involved inmetal binding.305 This study thus reveals a relationship between the structure, metal binding, andantibiotic activity of this antibiotic.
The structure of metal-free bacitracin has been determined by means of 2D-NMR spectroscopy, which revealed that the side chains of D-Phe9 and Ile8 are in close proximity of Leu3.307 However, theresult from the study of the Co2þ complex indicates that D-Phe9 and Ile5 are close to each other.305This difference is possibly induced by the metal binding. Bacitracin is known to bind to serineproteases, and the crystal structures of bacitracin–protease complexes have recently beendetermined.308 The protease-bound bacitracin has an extended structure, which prevents metal frombinding to the antibiotic. This structure of bacitracin is different from both the metal-free and metal-bound forms in solution determined by means of NMR. Bacitracin has also been known to inhibitmetalloproteases, presumably because of its metal-binding capability.309 In addition to proteaseinhibition, bacitracin can also inhibit a membrane-bound protein disulfide isomerase,310 and mayserve as a selective inhibitor of b1 and b7 integrin following a not yet known mechanism.311 Thus, the Figure 20. Proton NMR spectrum of the Co2þ complex of bacitracin in H2O at pH 5.5.The signals have been assigned as indicatedby the use of1D- and 2D-NMR techniques.
Figure 21. Stereo view of Co2 þ bacitracin A1produced by means of molecular modeling using nuclear magnetic relaxation ratesas distance constraints.The metal ion is coordinated to the drug through the nitrogen of thiazoline ring, the carboxylate of Glu4, andthe ring Ne of His10.The N-terminal amine is not bound to the metal, but may be hydrogen-bonded to Asn12.The side chains of Phe9and Ile5 are in close proximity and may serve as a flexible hydrophobic binding site for lipid pyrophosphates.This structure is expectedto be similar to the structures of the Co2 þ complexes of bacitracins B1and B2 with high antibiotic activities.
inhibitory property toward proteins may serve as a unique ‘‘alternative activity'' of this antibiotic,in addition to its better documented inhibition activity toward bacterial cell wall synthesis.
Ionophores3,312–316 and siderophores317 are relatively small molecules excreted by microorganismswhich can selectively bind and transport alkali or alkaline earth metal ions and Fe3þ , respectively,across cell membranes and artificial lipid bilayers. These molecules can serve as antibiotics by (a)disturbing the ionic balance across membrane via ion transport (particularly, the transport of alkaliand alkaline earth metal ions), such as nactins, lasalocid, and valinomycin, (b) creating pores onmembranes which results in leaking of cations through the pores, such as gramicidins, and (c)competing for essential iron in the environment, such as ferrichromes. The antibiotic activity ofionophores have also entitled them to be practically used as growth promoters and for increasingagricultural products.318 A. Structure, Cation Binding, and Transport of Ionophores Cation transport across the membrane by ionophores requires the participation of specific membraneproteins and is strictly regulated. The transport of cations results in disturbance of the ionic balanceacross the membrane upon release of the bound metal ions. This disturbance may slow down normalcell growth or even cause cell death. This family of cation-binding microbial products can thus beconsidered antibiotics. As opposed to the metalloantibiotics discussed in previous sections in whichthe metal ions serve as an integral part of the molecules to exhibit antibiotic activities, the metalions themselves in metalloionophores serve as the ‘‘magic bullets.'' The release of metal ions fromthe metalloionophores in the cells can cause imbalance of potential across cell membranewhich engenders antibiotic activities of ionophores. In the case of ion-channeling antibiotics, the‘‘magic bullets'' are transported directly into the cells to result in antibiotic effect. The mechanism ofthis type of antibiotic activity has been adopted in a recent design of channel-forming antibacterialagents.319 1. Structure and Metal Binding Metal binding to ionophores and siderophores is the key step that allows specific receptors on the cellsurface to recognize the metalloforms of the molecules, which results in transport of metal ions acrossthe membrane into the cell. Upon binding of metal ions, conformational changes of ionophores andsiderophores may occur.312 The structure of the metalloforms may vary dramatically, depending onthe bound metal ions. In the case of enniatin (which has a cyclic[L-N-methyl-valine-D-hydroxy-isovalerate]3 structure), the parent ionophore has a structure very similar to its K þ complex, whereasthe Rb2–enniatin complex has a structure quite distinct from that of the metal-free form.320 Thisdifference is attributed to the different ionic radii and binding affinities of alkali metal ions withenneatin. In many other cases, the binding of metal ions results in significant changes of the struc-tures of ionophores from more extended conformations to more folded forms on the basis of thecrystal structures of nactins (e.g., nonactin, tetranactin, and dinactin) and valinomycin and theirmetalloforms.321,322 A common structural feature of this family of antibiotics is the presence of an O-rich metal binding environment, including ether and ketide linkages, the carbonyl group of esters and amides,and carboxylates. Schematic structures of a few ionophores are shown in Figure 22. Such polyketidestructure is synthesized by polyketide synthases via C–C bond (type I and II polyketide synthases) andC–O bond (type III polyketide synthase) formation, analogous to the peptide synthetase for bacitracinsynthesis (Section 4) and peptide/ketide synthetase for BLM synthesis (Section 2.A ‘‘Bleomycin'').
Type III polyketide synthase has been demonstrated to be involved in the synthesis of the C–Oformation in the antibiotic nonactin (Fig. 22A).75m Since these O-rich moieties are preferred ligands for alkali and alkaline earth metal ions, the preference in metal binding is thus not because of the ligands but controlled by different mechanisms Figure 22. Schematic structures of nactins (A): Nonactin (R1, R2, R3, and R4 ¼ CH3), monactin (R1 ¼ C2H5; R2, R3, and R4 ¼ CH3),trinactin (R1and R3 ¼ C2H5; R2 and R4 ¼ CH3), and tetranactin (R1, R2, R3, and R4 ¼ C2H5); valinomycin, which has a cyclic struc-ture with three repeating units of (L-Val L-Lactate D-Val D-hydroxylisovalerate) (B); lasalocid (C), and monensin (D).
such as the different molecular structures, the different sizes of the metal-binding site, the differentionic radii of metal ions (i.e., 0.66, 0.95, 1.33, 1.48, and 1.69 A ˚ for Liþ , Naþ , Kþ , Rbþ , and Csþ , respectively), the use of different moieties for metal binding, and/or the different degrees of hydrationenergy of cations. For instance: (i) while lasalocid with a linear structure has a larger apparent affinityfor its Ca2þ binding than its Kþ binding in the presence of vesicle membrane,323 the cyclic nactinsexhibit higher preference toward monovalent cation binding and best in the binding with NH þ The larger binding cavity in nonactin than in tetranactin (Fig. 22A) results in a binding affinity ofCsþ > Naþ for nonactin, but Naþ > Csþ for tetranactin.324 (iii) The large ‘‘metal binding cavity'' invalinomycin (Fig. 22B) allows its binding with one (e.g., K þ ) or two (e.g., Ba2þ ) metal ions.312 (iv)X-ray crystal structures show that Kþ is bound to valinomycin at the ‘‘internal binding site,'' whereasNa þ at an ‘‘external binding site.''322 (v) While the apparent formation constants for Kþ and Naþbinding to nonactin in dry acetone are in the same order, that for Na2þ is significantly decreased in thepresence of water which indicates the importance of hydration in cation binding to nactins.325 2. Carboxylic Ionophores This family of ionophores are comprised of a linear polycycloether backbone and a carboxylate group,as represented by lasalocid and monensin (structures C and D, Fig. 22). Most of these ionophores bindmonovalent cations with a one-to-one ratio and bind divalent cations with a metal:ligand ratio of 1:2,in which the cyclic ether-O atoms serve as metal-binding ligand.326 Recent FT-IR and 7Li- and 23Na-NMR studies suggest that lasalocid forms a fluxional 1:1 complex with Li þ and 2:2 complexes withKþ , Rbþ , and Csþ ions in solid and in chloroform,327 and 1:1 and 2:2 complexes with Naþ inchloroform.328 The neutral and acid complexes of lasalocid with alkali metals, Tlþ , Agþ , andalkaline earth metal ions have been studied with 1H- and 13C-NMR, and have been suggested topossess similar structures.329 Lanthanide(III) (Ln3þ ) complexes of lasalocid with different metal-to-drug ratios have also been reported.330 Extraction of water-soluble Ln3þ–acetylacetonato (acac)complexes into organic solvent by lasalocid was observed, in which ternary lasalocid–Ln3þ–acaccomplexes are proposed to form with high preference toward smaller Ln3þ ions.331 Metal com-plexation of lasalocid has recently been reviewed.332 3. Gramicidin Family Gramicidins are a family of peptide ionophores3,333 produced by Bacillus brevis as a mixture of a fewcongeners with different amino acid compositions, of which Gramicidin A is the major componentwhose primary structure is shown below.3 Formyl-L-Val1– Gly2– L-Ala3– D-Leu4– L-Ala5– D-Val6– L-Val7– D-Val8– L-Trp9– D-Leu10– L-Trp11– D-Leu12– L-Trp13– D-Leu14– L-Trp15– ethanolamine.
L-Trp11 in gramicidin A is replaced by L-Phe in gramicidin B, and by L-Tyr in gramicidin C. This family of ion channeling antibiotics exhibit a different mechanism for their action from the cation-carrying ionophores described in the above section. Gramicidins can insert into the lipid bilayer as adimer and span across the membrane, forming a unique b-double-stranded helix which creates achannel of 4 A ˚ wide for cations to permeate (Fig. 23).334 This channel exhibits an cation selectivity of H þ > NHþ > Csþ > Rbþ > Kþ > Naþ > Liþ > N(CH3)4 in 0.1 M salt,335 yet does not show permeability to divalent metal ions Mg2þ , Ca2þ , Ba2þ , and Zn2þ . The divalent metal ions canbind to the b-double-stranded helix structure at the entrance of the channel which prevents thetransport of monovalent cations.336 The trend in the alkali metal binding affinity follows the sameorder as the relative ionic mobility, ionic radius, and hydration energy of the ions.3 The structures of several gramicidin congeners and their metal-bound forms have been de- termined with X-ray crystallography337 and NMR spectroscopy in both micelles338 and solidstates333,339,340 (Fig. 23). Similar configurations are observed for the peptide backbone in bothsolution and solid states.337–341 However, deviations are observed among side chains, particularly Figure 23. Structures of gramicidin A determined with NMR spectroscopy (top, Protein Data Bank ID: 1MIC.pdb) and X-ray crystal-lography (middle, top view; bottom, side view, Protein Data Bank ID: 1AV2).Three Cs2þ ions located in the channel are found in thecrystal structure.
Trp-9. The structures in solution determined with NMR spectroscopy exhibit a higher degree ofirregularity than the structures in solid state determined with crystallography, such as the shape of thechannel (Fig. 23, top). The binding of monovalent metal ions does not seem to cause significantstructural change of gramicidins. The structure similarity of this antibiotic with and without boundcations is quite reasonable since the insertion of gramicidins into membranes does not rely on thebinding of metal ions as in the case of other ionophores described above. In addition, the similaritydoes not create further energy barrier for metal binding and transportation. According to thesestructures, the hydrophobic amino acid side chains are located outside the channel. This moleculararrangement allows gramicidins to exhibit extensive hydrophobic interaction with membrane uponinsertion into the membrane, whereas the carbonyl groups are positioned inside the channel forinteraction with and transportation of cations.
The structures of gramicidins A, B, and C in micelles obtained by means of 2D-NMR techniques are very similar, with background atom RMSD 90.5 A ˚ ascribed to similar structures.338 The side chains in these three congeners also have similar configurations. Despite their very similar structuresand monovalent cation specificity, gramicidins A, B, and C exhibit different cation binding andtransporting properties, such as the conductance and activation energy for ion transport. The minorvariations in the structures of the three congeners cannot explain these significantly differentproperties. Change in dipole moment of the side chains (i.e., dipole moments of 2.08, 0, and 1.54 D forTrp, Phe, and Tyr, respectively) was suggested to cause the difference. The decrease in the singlechannel conductance of Naþ for several natural and synthetic gramicidin congeners was found tocorrelate with the replacement of a Trp by Phe, which decreases the dipole moment.338a B. Iron Sequestering and Antibiotic Activities The extremely small solubility product Ksp of 1038 for Fe(OH)3 makes soluble Fe3þ in aqueoussolutions very scarce. To overcome this low availability of iron under aerobic conditions,microorganisms excrete Fe3þ -specific siderophores (Fig. 24) which bind Fe3þ with extraordinarilyhigh affinity constants in the range of 1030–1052 M1 and transport Fe3þ into cells via specificreceptors.317,342 Once the Fe3þ complexes of siderophores are transported into the cells, the iron canbe released upon reduction. In addition to the iron transport activity of siderophores, the Cu2þ , Co3þ ,and Ni2þ complexes of the siderophore desferal were found to exhibit an interesting ‘‘alternativebioactivity'' toward the cleavage of plasmid DNA and oligonucleotides343 which points a newdirection for design of new ‘‘chemical nucleases.'' 1. Mechanism and Structure of Siderophores The biosynthesis of siderophores is regulated by cellular concentration of Fe2þ .344 When theconcentration is low, Fe2þ is dissociated from an ferric uptake regulatory (Fur) protein, resulting inthe binding of Fur to the operon that initiates the synthesis and excretion of siderophores for Fe3þsequestering.342 The Fe3þ–siderophore complexes can be recognized by species-specific receptorsand transported into the cells. For example, although ferrichrome A serves as an iron carrier for fungi,it does not serve that purpose in bacteria; whereas ferrichrome does.345 Thus, a rational approach tochemotherapy becomes possible when it is based on iron transport mechanism in microorganisms317 Figure 24. Two prototypical families of Fe-free siderophores: (A) hydroxamate-based ferrioxamines B (R ¼ NHþ (R ¼ NHCOCH3), and E (R ¼ NHCOCH2CH2 thatclosesthe chainto give a 3-fold symmetric ringstructure) and (B) catechol-basedenterobactin.
by (a) controlling iron transport via iron chelating, (b) transport of an antibiotic substance viaconjugating with siderophores (e.g., albomycin discussed below), and/or (c) inactivation of sidero-phore receptors via inhibitor binding.
The crystal structures of several hydroxamate-containing Fe3þ -bound siderophores have been resolved which allow a detailed comparison of their structures and correlation of their structures withfunction. While the crystal structures show that the iron complexes of ferrioxamines B,346 D1,347 andE348 and desferrioxamine E349 are mixture of L- and D-cis isomers, ferrichrome complexes350 areshown to be exclusively the L-cis isomers. This cis configuration seems to be important for therecognition of these siderophores by membrane receptors (see next Section), wherein the ‘‘carbonylface'' is considered to play an important role since the ‘‘oxime face'' is relatively more shielded thanthe carbonyl face in the structures of these siderophores.
2. Albomycin Structure and Receptor Binding Albomycin is a prototypical siderophore antibiotic produced by Streptomyces. It is a broad-spectrumantibiotic active against several Gram-positive and -negative bacteria with low minimum inhibitionconcentrations, and is even active toward those bacteria resistant to penicillin, streptomycin, TC, anderythromycin.317,351 This antibiotic is a natural conjugate comprised of an iron-binding tri-d-N-hydroxy-d-N-acetyl-L-ornithine site analogous to that of ferrichromes and an antibiotic moiety ofthioribosyl pyrimidine (Fig. 25). Albomycin is recognized by the ferrichrome receptor in the outercell membrane of E. coli,352 and is activated after cleavage by peptidase N to release the antibioticmoiety. Thus, albomycin-resistant bacteria are found to lack either the receptor353 or the peptidase.354 Iron-depleted siderophores are expected to have quite flexible and extended structures. Upon Fe3þ binding, the metal binding ligands are held together by the metal and exhibits a compact metal-binding configuration which is recognized by the ferrichrome receptors (cf. Fig. 25). A few structuresof the membrane transporter protein FhuA (ferric hydroxamate uptake A protein) with and without abound Fe3þ–sederophore have been determined (Fig. 25).355 The Fe3þ binding moiety of the Fe3þcomplexes of phenylferricrocin, albomycin, ferrichrome, and ferricrocin have quite similar coordina-tion chemistry upon their binding to FhuA, and are bound to the receptor in a similar orientation andinteracts with the same amino acid side chains, including Tyr and Arg side chains. The cis con-figurations of the Fe3þ–siderophore complexes bound to the uptake proteins are found to be the sameas those of Fe3þ–siderophores without bound to the protein.346–350 Similar interactions and confi-guration are also observed in the crystal structure of gallichrome–FhuD complex.356 The binding ofFe3þ–ferichrome to FhuA induces a significant conformational change through the N-terminaldomain, including an unwinding of helix 2 near the binding site and a 11-A ˚ translation of the loop next to it, as well as a 17-A ˚ moving of Trp22 on the oppose side from the binding site. How these changes result in the uptake of the Fe3þ–siderophore complexes and transport through cell membranecould not be revealed in the crystallographic studies.
Recently, a semi-synthetic rifamycin analogue CGP4832 was found to bind to and actively transported by FhuA protein, despite its distinctly different structure from those of albomycin andferrichrome.357 In addition to FhuA, the membrane transporters FepA (ferric enterobactin transportprotein)358 and FecA (ferric citrate uptake protein which recognizes and transports the dinuclearFe(III)2–citrate2 complex)359 are also characterized to contain the 22-stranded b-barrel structure asfound in FhuA,360 with a cross section of 35–45 A ˚ . The above studies have provided the molecular basis for rational design of antibiotic siderophores that target bacteria through specific siderophorereceptors of the bacteria.
C. Perspectives of Metal Ions in Medicine In addition to the metalloantibiotics discussed in this review, a number of drugs and potentialpharmaceutical agents also contain metal-binding or metal-recognition sites, which can bind or


Figure 25. Top: Schematic structure of albomycin. Middle: The crystal structure of the Fe3þ albomycin FhuA complex (ProteinData Bank ID1QKC).The Fe3þ drug complex is shown in red color (Fe3þ in green), and the protein as gray ribbons.The binding siteof the drug complex is located inside the FhuA‘pocket.' Bottom: The structure of Fe3þ albomycin seen in the crystal structures ofthe Fe3þ albomycin FhuA complex, in which the residues involved in H-bonding with the Fe33þ albomycin are shown inblue color.
interact with metal ions and potentially influence their bioactivities and might also cause damages ontheir target biomolecules. Numerous examples these ‘‘metallodrugs'' and ‘‘metallopharmaceuticals''and their actions can be found in the literature, for instance: (a) several anti-inflammatory drugs, suchas aspirin and its metabolite salicylglycine,361 ibuprofen,362 the indole derivative indomethacin,363bioflavonoid rutin,364 diclofenac,365 suprofen,366 and others367 are known to bind metal ions andaffect their antioxidant and anti-inflammatory activities; (b) the potent histamine-H2-receptorantagonist cimetidine368 can form complexes with Cu2þ and Fe3þ , and the histidine H2 blocker antiulcer drug famotidine can also form stable complex with Cu2þ ;369 (c) the anthelmintic andfungistatic agent thiabendazole, which is used for the treatment of several parasitic diseases, forms aCo2þ complex with metal:drug ratio of 1:2;370 (d) the Ru2þ complex of the anti-malaria agentchloroquine exhibits an activity two to five times higher than the parent drug against drug-resistantstrains of Plasmodium faciparum;371 (e) a number of Ru2þ /3þ and Rh2þ /3þ complexes are found tobind DNA and exhibit anti-tumor activities;256,372 (f) the antiviral trifluoperazine forms complexeswith VO2þ , Ni2þ , Cu2þ , Pd2þ , and Sn4 þ which exhibit higher inhibition activities than the metal-free drug when tested on Moloney murine leukemia virus reverse transcriptase;373 (g) the clinicallyuseful b-lactamase inhibitor sulbactam can form complexes with Ni2þ , Cu2þ , and Fe3þ ;374 (h) a fewhormone-anchored metallodrugs have been prepared which show enhanced receptor binding andhigher activities against cancer cells;241,375 (i) the thiosemicarbazone-conjugated isatin (which showsa broad-spectrum bioactivity376) can bind late first-row transition metal ions and exhibit activitytoward human leukemia cell lines, however, without inducing cell apoptosis;377 and (j) metalcomplexes (including Be2þ , Mg2þ , Mn2þ , Co2þ , Ni2þ , Cu2þ , Zn2þ , Cd2þ , Pb2þ , Fe3þ , Al3þ , andLa3þ ) of several carbonic anhydrase inhibiting sulfonamides378 have been investigated for theirtopical intraocular pressure lowering properties and as potential agents against gastric acidimbalances.
There are also a number of metallodrugs and metallopharmaceuticals which have been utilized for the treatment of diseases and disorders or as diagnostic agents,233e,379,380 such as gold antiarthriticdrugs, bismuth antiulcer drugs, gadolinium MRI contrast agents, technetium radiopharmaceuticals,metal-based X-ray contrast agents, and photo- and radio-sensitizers, vanadium as insulin mimics, andlithium psychiatric drugs. The metal ion Liþ can be considered the smallest effective metallodrugwhose carbonate and citrate salts exhibit significant therapeutic benefit in the treatment of manicdepression (bipolar mood disorder).381 Some recent studies by means of 3D-MRI techniques indicatethat the volume of the brain gray matter is increased in bipolar disorder patients treated with Liþ .382The status of Liþ in cells have been extensively studied and recently reviewed.381,383 It is alsointeresting to point out that the metal ion Sb3þ may be regarded as the simplest ‘‘metalloantibiotic''from a broader viewpoint of the term, whose salts (including N-methylglucamine antimonite and Na-stibogluconate) have been utilized for the treatment of leishmaniasis against the protozoan parasiteLeishmania.384 The antiprotozoal mechanism of Sb3þ is thought to be attributed to its binding totrypanothione that is essential for the growth of the parasite.
This review has summarized the structure, function, and activity of several different families of metalloantibiotics, and has also pointed out the design and potential utilization of metal complexes forbattling pathogenic microorganisms. Because of the increase in bacterial resistance toward manycurrently used antibiotics in recent years, further development of new antibiotics has become anurgent mission. Better understanding of the structure, function, and mechanism of existingmetalloantibiotics and the mechanism of antibiotic resistance will lead to better design of metalcomplexes for this mission. As the chemical properties of metal ions can vary significantly and canalso be further fine-tuned by proper design of drug ligands for targeting different biomolecules andbiocomponents, new generations of various ‘‘metalloantibiotics'' isolated from natural resources orobtained via chemical syntheses and/or modifications that exhibit more effective antiparasitic,antibacterial, antiviral, and anti-tumor activities can be foreseen.
The NMR studies of bleomycin, anthracyclines, streptonigrin, aureolic acids, and bacitracin carriedout in the author's laboratory have been supported by the start-up funds, Research and CreativeScholarship Grants and the PYF Award of the University of South Florida, and by the Edward L.
Cole Research Grant (F94USF-3) of the American Cancer Society—Florida Division. The author's co-workers Dr. Xiangdong Wei, Dr. Jon Epperson, and Dr. Jason Palcic have made significantcontributions in better understanding of structure–function relationship of several metalloantibioticsdiscussed in this review.
1. David SA, Balasubramanian KA, Mathan VI, Balaram P. Analysis of the binding of polymyxin B to endotoxic lipid A and core glycolipid using a fluorescent displacement probe. Biochim Biophys Acta1992;1165:145–152.
2. (a) Lown JW. Discovery and development of anthracycline antitumor antibiotics. Chem Soc Rev 1993;22:165–176. (b) Phillips DR, Cullinane C, Trist H, White RJ. In vitro transcription analysis of thesequence specificity of reversible and irreversible complexes of adriamycin with DNA. In: Pullman B,Jortner J, editors. Molecular basis of specificity in nucleic acid–drug interactions. Netherlands: Kluwer;1990. pp. 137–155.
3. Hinton JF, Koeppe RE II. Complexing properties of gramicidins. Metal Ions Biol Syst 1985;19:173–206.
4. Parker MW, Pattus F, Tucher AD, Tsernoglou D. Structure of the membrane-pore-forming fragment of colicin-A. Nature 1989;337:93–96.
5. (a) Oka T, Hashizume K, Fujita H. Inhibition of peptidoglycan transpeptidase by beta-lactam antibiotics: Structure–activity relationships. J Antibiot 1980;33:1357–1362. (b) Fre re JM, Ghuysen JM, Reynolds PE,Moreno R. Binding of beta-lactam antibiotics to the exocellular DD-carboxypeptidase-transpeptidase ofStreptomyces R39. Biochem J 1974;143:241–249.
6. (a) Umezawa H, Ishizuka M, Suda H, Hamadam M, Takeuchi T. Bestatin, an inhibitor of aminopeptidase B, produced by actinomyces. J Antibiot 1976;29:97–99. (b) Umezawa H, Ishizuka M, Aoyagi T, Takeuchi T.
Enhancement of delayed-type hypersensitivity by bestatin, an inhibitor of aminopeptidase B and leucineaminopeptidase. J Antibiot 1976;29:857–859. (c) Osada T, Ikegami S, Takiguchi-Hayashi K, Yamazaki Y,Katoh-Fukui Y, Higashinakagawa T, Sakaki Y, Takeuchi T. Increased anxiety and impaired pain response inpuromycin-sensitive aminopeptidase gene-deficient mice obtained by a mouse gene-trap method.
J Neurosci 1999;19:6068–6078.
7. Waksman SA. Streptomycin: Background, isolation, properties, and utilization. Nobel Lecture, December 8. For physical methods, see: Que L, Jr., editor. Physical methods in bioinorganic chemistry, spectroscopy, and magnetism. University Science Books; 2000.
9. Guschlbauer W, Saenger W, editors. DNA–ligand interactions: From drugs to proteins (NATO ASI Series.
Series A, Life Science, Vol. 137), New York: Plenum; 1987. (b) Neville RK, editor. Chemistry and physicsof DNA–ligand interactions. New York: Adenine Press; 1990. (c) Chaires JB, Waring MJ. Drug–nucleicacid interactions. Met Enzymol Vol. 340, San Diego, CA: Academic Press; 2001.
10. Tullius TD, editor. Metal-DNA Chemistry. ACS Symposium Series 402, the American Chemical Society, 1989. (b) Barton JK. Metal/nucleic acid interactions. In: Bertini I, Gray HB, Lippard SJ, Valentine JS,editors. Bioinorganic Chemistry. University Science Books; 1994.
11. Umezawa H, Maeda K, Takeuchi T, Okami Y. New antibiotics, bleomycin A and B. J Antibiot 1966;19: 12. Takeshita M, Horwitz SB, Grollman AP. Bleomycin, an inhibitor of vaccinia virus replication. Virology 1974;60:455–456. (b) Takeshita M, Grollman AP, Horwitz SB. Effect of ATP and other nucleotides on thebleomycin-induced degradation of vaccinia virus DNA. Virology 1976;69:453–463. (c) Takeshita M,Horwitz SB, Grollman AP. Mechanism of the antiviral action of bleomycin. Ann NY Acad Sci 1977;284:367–374.
13. Lazo JS, Sebti SM, Schellens JH. Bleomycin. Cancer Chemother Biol Resp Modif 1996;16:39–47.
14. (a) Burger RM. Nature of activated bleomycin. Struct Bond 2000;97:287–303. (b) Burger RM. Cleavage of nucleic acids by bleomycin. Chem Rev 1998;98:1153–1170. (c) Claussen CA, Long EC. Nucleic acidrecognition by metal complexes of bleomycin. Chem Rev 1999;99:2797–2816. (d) Boger DL, Cai H.
Bleomycin: Synthetic and mechanistic studies. Angew Chem Int Ed 1999;38:449–476. (e) Stabbe J,Kozarich JW. Mechanism of bleomycin-induced DNA degradation. Chem Rev 1987;87:1107–1136. (f)Petering DH, Byrnes RW, Antholine WE. The role of redox-active metals in the mechanism of action ofbleomycin. Chem Biol Inter 1990;73:133–182. (g) Hecht SM. The chemistry of activated bleomycin. AccChem Res 1986;19:383–391.
15. Umezawa H, Takita T. The bleomycins: Antitumor copper-binding antibiotics. Struct Bond 1980;40: 16. Kane SA, Hecht SM. Polynucleotide recognition and degradation by bleomycin. Prog Nucl Acids Res Mol 17. (a) Takeshita M, Grollman AP, Ohtsubo E, Ohtsubo H. Interaction of bleomycin with DNA. Proc Nat Acad Sci USA 1978;75:5983–5987. (b) Cullinam EB, Gawron LS, Rustum YM, Beerman TA. Extrachromo-somal chromatin: Novel target for bleomycin cleavage in cells and solid tumors. Biochemistry1991;30:3055–3061.
18. (a) Chang C-H, Meares CF. Cobalt-bleomycins and deoxyribonucleic acid: Sequence-dependent interactions, action spectrum for nicking, and indifference to oxygen. Biochemistry 1984;23:2268–2274. (b) Saito I, Morii T, Sugiyama H, Matsura T, Meares CF, Hecht SM. Photoinduced DNA strandscission by cobalt bleomycin green complexes. J Am Chem Soc 1989;111:2307–2308. (c) Worth L, Jr.,Frank BL, Christner DF, Absalon MJ, Stubbe J, Kozarich JW. Isotope effects on the cleavage of DNA bybleomycin: Mechanism and modulation. Biochemistry 1993;32:2601–2609.
19. (a) Absalon MJ, Stubbe J, Kozarich JW. Sequence-specific double-strand cleavage of DNA by Fe- bleomycin. 1. The detection of sequence-specific double-strand breaks using hairpin oligonucleotides.
Biochemistry 1995;34:2065–2075. (b) Absalon MJ, Wu W, Stubbe J, Kozarich JW. Sequence-specificdouble-strand cleavage of DNA by Fe-bleomycin. 2. Mechanism and dynamics. Biochemistry1995;34:2076–2086.
20. Sugiyama H, Kilkuskie RE, Chang L-H, Ma L-T, Hecht SM, van der Marel GA, van Boom JH. DNA strand scission by bleomycin—Catalytic cleavage and strand selectivity. J Am Chem Soc 1986;108:3852–3854.
21. Chakrabarti S, Makrigiorgos GM, O'Brien K, Bump E, Kassis AI. Measurement of hydroxyl radicals catalyzed in the immediate vicinity of DNA by metal–bleomycin complexes. Free Rad Biol Med1996;20:777–783.
22. Povirk LF, Austin MJF. Genotoxicity of bleomycin. Mut Res 1991;257:127–143.
23. (a) Hecht SM. RNA degradation by bleomycin, a naturally occurring bioconjugate. Bioconj Chem 1994;5:513–526. (b) Mascharak P, Hu¨ttenhofer A. Cleavage of RNA by Fe(II)-bleomycin. In: Schroeder R,Wallis MG, editors. RNA-binding antibiotics. Landes: BioScience; 2001.
24. (a) Magliozzo RS, Peisach J, Ciriolo MR. Transfer RNA is cleaved by activated bleomycin. Mol Pharmocol 1989;35:428–432. (b) Carter BJ, de Vroom E, Long EC, van der Marel GA, van Boom JH, Hecht SM. Site-specific cleavage of RNA by Fe(II)-bleomycin. Proc Nat Acad Sci USA 1990;87:9373–9377. (c)Hu¨ttenhofer A, Hudson S, Noller HF, Mascharak PK. Cleavage of tRNA by Fe(II)-bleomycin. J Biol Chem1992;267:24471–24475.
25. (a) Carter BJ, Reddy KS, Hecht SM. Polynucleotide recognition and strand scission by Fe-bleomycin.
Tetrahedron 1991;47:2463–2474. (b) Holmes CE, Carter BJ, Hecht SM. Characterization of iron(II)-bleomycin-mediated RNA stand scission. Biochemistry 1993;32:4293–4307.
26. (a) Morgan MA, Hecht SM. Iron(II) bleomycin-mediated degradation of a DNA–RNA hetero-duplex.
Biochemistry 1994;33:10286–10293. (b) Bansal M, Lee JS, Stubbe JA, Kozarich JW. Mechanisticanalyses of site-specific degradation in DNA–RNA hybrids by prototypic DNA cleavers. Nucleic AcidsRes 1997;25:1836–1845.
27. Dabrowiak JC. The coordination chemistry of bleomycin: A review. J Inorg Biochem 1980;13:317–337.
28. (a) Burger RM, Freedman JH, Horwitz SB, Peisach J. DNA-degradation by manganese(II)-bleomycin plus peroxide. Inorg Chem 1984;23:2215–2217. (b) Ehrenfeld GM, Murugesan N, Hecht SM. Activation ofoxygen and mediation of DNA-degradation by manganese bleomycin. Inorg Chem 1984;23:1496–1498.
29. (a) Ishida R, Takahashi T. Increased DNA chain breakage by combined action of bleomycin and superoxide radical. Biochem Biophys Res Commun 1975;66:1432–1438. (b) Sausville EA, Stein RW, Peisach J,Horeitz SB. Properties and products of degradation of DNA by bleomycin and Iron(II). Biochemistry1978;17:2746–2754. (c) Absalon MJ, Kozarich JW, Stubbe J. Sequence-specific double-strand cleavage ofDNA by Fe-bleomycin. 1. The detection of sequence-specific double-strand breaks using hairpinoligonucleotides. Biochemistry 1995;34:2065–2075. (d) Absalon MJ, Wu W, Kozarich JW, Stubbe J.
Sequence-specific double-strand cleavage of DNA by Fe-bleomycin. 2. Mechanism and dynamics.
Biochemistry 1995;34:2076–2086. (e) McGall GH, Rabow LE, Ashley GW, Wu SH, Kozarich JW, StubbeJ. New insight into the mechanism of base propenal formation during bleomycin-mediated DNAdegradation. J Am Chem Soc 1992;114:4958–4967.
30. Sugiura Y. Monomeric cobalt(II)-oxygen adducts of bleomycin antibiotics in aqueous solution—A new ligand type for oxygen binding and effect of axial Lewis base. J Am Chem Soc 1980;102:5216–5221.
31. Greenaway FT, Dabrowiak JC, Van Husen M, Grulich R, Crooke ST. The transition metal binding properties of a 3rd generation bleomycin analogue, tallysomycin. Biochem Biophys Res Commun1978;85:1407–1414.
32. (a) Sugiura Y. Metal coordination core of bleomycin: Comparison of metal complexes between bleomycin and its biosynthetic intermediate. Biochem Biophys Res Commun 1979;87:643–648. (b) Guan LL, Totsuka R, Kuwahara J, Otsuka M, Sugiura Y. Cleavage of yeast transfer RNAphe with Ni(III) and Co(III)complexes of bleomycin. Biochem Biophys Res Commun 1993;191:1338–1346. (c) Sugiura Y, Mino Y.
Nickel(III) complexes of histidine-containing tripeptide and bleomycin–Electron-spin resonancecharacteristics and effect of axial nitrogen donors. Inorg Chem 1979;18:1336–1339. (d) Guan LL,Kuwahara J, Sugiura Y. Guan LL, Kuwahara J, Sugiura Y. Guanine-specific binding by bleomycinnickel(III) complex and its reactivity for guanine-quartet telomeric DNA. Biochemistry 1993;32:6141–6145.
33. (a) Dabrowiak JC, Greenaway FT, Grulich R. Transition-metal binding site of bleomycin A2. A carbon-13 nuclear magnetic resonance study of the zinc(II) and copper(II) derivatives. Biochemistry 1978;17:4090–4096. (b) Dabrowiak JC, Greenaway FT, Longo WE, Van Husen M, Crooke ST. A spectroscopicinvestigation of the metal binding site of bleomycin A2. The Cu(II) and Zn(II) derivatives. BiochimBiophys Acta 1978;517:517–526. (c) Sugiura Y. The production of hydroxyl radical from copper(I)complex systems of bleomycin and tallysomycin: Comparison with copper(II) and iron(II) systems.
Biochem Biophys Res Commun 1979;90:375–383.
34. Ehrenfeld GM, Shipley JB, Heimbrook DC, Sugiyama H, Long EC, van Boom JH, van der Marel GA, Oppenheimer NJ, Hecht SM. Copper-dependent cleavage of DNA by bleomycin. Biochemistry1987;26:931–942.
35. Otvos JD, Antholine WE, Wehrl S, Petering DH. Metal coordination environment and dynamics in cadmium-113 bleomycin: Relationship to zinc bleomycin. Biochemistry 1996;35:1458–1465.
36. Papakyriakou A, Mouzopoulou B, Katsaros N. The detailed structural characterization of the Ga(III)– bleomycin A2 complex by NMR and molecular modeling. J Inorg Biochem 2001;86:371 (ICBIC10abstract).
37. Subramanian R, Meares CF. Photo-induced nicking of deoxyribonucleic acid by ruthenium(II)-bleomycin in the presence of air. Biochem Biophys Res Commun 1985;133:1145–1151.
38. Brooks RC, Carnochan P, Vollano JF, Powell NA, Zweit J, Sosabowski JK, Martellucci S, Darkes MC, Fricker SP, Murrer BA. Metal complexes of bleomycin: Evaluation of [Rh-105]-bleomycin for use intargeted radiotherapy. Nucleic Med Biol 1999;26:421–430.
39. Lever ABP. Inorganic electronic spectroscopy, 2nd edn. Amsterdam: Elsevier; 1984.
40. Itaka Y, Nakamura H, Nakatani T, Muraoka Y, Fujii A, Takita T, Umezawa H. Chemistry of bleomycin. XX.
The X-ray structure determination of P-3A Cu(II)-complex a biosynthetic intermediate of bleomycin.
J Antibiot 1978;31:1070–1072.
41. Burger RM, Adler AD, Horwitz SB, Mims WB, Peisach J. Demonstration of nitrogen coordination in metal–bleomycin complexes by electron spin-echo envelope spectroscopy. Biochemistry 1981;20:1701–1704.
42. (a) Loeb KE, Zaleski JM, Westre TE, Guajardo RJ, Mascharak PK, Hedman B, Hodgson KO, Solomon EI.
Spectroscopic dedinition of the geometric and electronic structure of the nonheme active site in iron(II)bleomycin–correlation with oxygen reactivity. J Am Chem Soc 1995;117:4545–4561. (b) Loeb KE,Zaleski JM, Hess CD, Hecht SM, Solomon EI. Spectroscopic investigation of the metal ligation andreactivity of the ferrous active sites of bleomycin and bleomycin derivatives. J Am Chem Soc 1998;120:1249–1259.
43. (a) Oppenheimer NJ, Chang C, Chang LH, Ehrenfeld G, Rodriguez LO, Hecht SM. Deglyco-bleomycin.
Degradation of DNA and formation of a structurally unique Fe(II)  CO complex. J Biol Chem1982;257:1606–1609. (b) Akkerman MAJ, Neijman EWJF, Wijmenga SS, Hilbers CW, Bermel W.
Studies of the solution structure of the bleomycin A2 iron(II) carbon monoxide complex by means of2-dimensional NMR spectroscopy and distance geometry calculations. J Am Chem Soc 1990;112:7462–7474.
44. (a) Akkerman MAJ, Haasnoot CA, Hilbers CW. Studies of the solution structure of the bleomycin-A2-zinc complex by means of two-dimensional NMR spectroscopy and distance geometry calculations. Eur JBiochem 1988;173:211–225. (b) Akkerman MAJ, Haasnoot CAG, Pandit UK, Hilbers CW. Completeassignment of the C-13 NMR spectra of bleomycin A2 and its zinc complex by means of two-dimensionalNMR spectroscopy. Magn Reson Chem 1988;26:793–802. (c) Williamson D, McLennan IJ, Bax A,Gamcsik MP, Glickson JD. Two-dimensional NMR study of bleomycin and its zinc(II) complex:Reassignment of 13C resonances. J Biomol Struct Dyn 1990;8:375–398.
45. Calafat A. Won H, Marzilli LG. A new arrangement for the anticancer antibiotics tallysomycin and bleomycin when bound to zinc: An assessment of metal and ligand chirality by NMR and moleculardynamics. J Am Chem Soc 1997;119:3656–3664.
46. (a) Sugiura Y. Monomeric cobalt(II)-oxygen adducts of bleomycin antibiotics in aqueous solution. A new ligand type for oxygen biding and effect of axial Lewis base. J Am Chem Soc 1980;102:5216–5221. (b)Sugiura Y. Oxygen binding to cobalt(II)-bleomycin. J Antibiot 1978;31:1206–1208.
47. Palmer G. Electron paramagnetic resonance of metalloproteins. In: Que L, Jr., editor. Physical methods in bioinorganic chemistry, spectroscopy, and magnetism. Sausalito, CA: University Science Books; 2000.
48. (a) Xu RX, Antholine WE, Petering DH. Reaction of Co(II)bleomycin with dioxygen. J Biol Chem 1992;267:944–949. (b) Xu RX, Antholine WE, Petering DH. Reaction of DNA-bound Co(II)bleomycinwith dioxygen. J Biol Chem 1992;267:950–955.
49. Caceres-Cortes J, Sugiyama H, Ikudome K, Saito I, Wang AH. Interactions of deglycosylated cobalt(III)- pepleomycin (green form) with DNA based on NMR structural studies. Biochemistry 1997;36:9995–10005.
50. Wu W, Vanderwall DE, Lui SM, Tang XJ, Turner CJ, Kozarich JW, Stubbe J. Studies of Co center dot Bleomycin A2 green: Its detailed structural characterization by NMR and molecular modeling and itssequence-specific interaction with DNA oligonucleotides. J Am Chem Soc 1996;118:1268–1280.
51. (a) Vanderwall DE, Lui SM, Wu W, Turner CJ, Kozarich JW, Stubbe J. A model of the structure of HOO-Co-bleomycin bound to d(CCAGTACTGG): Recognition at the d(GpT) site and implications fordouble-stranded DNA cleavage. Chem Biol 1997;4:373–387. (b) Hoehn ST, Junker HD, Bunt RC, TurnerCJ, Stubbe J. Solution structure of Co(III)-bleomycin-OOH bound to a phosphoglycolate lesion containingoligonucleotide: Implications for bleomycin-induced double-strand DNA cleavage. Biochemistry2001;40:5894–5905. (c) Wu W, Vanderwall DE, Teramoto S, Lui SM, Hoehn ST, Tang XJ, Turner CJ,Boger DL, Kozarich JW, Stubbe J. NMR studies of Co center dot deglycoBleomycin A2 green and itscomplex with d(CCAGGCCTGG). J Am Chem Soc 1998;120:2239–2250. (d) Lui SM, Vanderwall DE,Wu W, Tang XJ, Turner CJ, Kozarich JW, Stubbe J. Structural characterization of Co center dot bleomycinA2 brown: Free and bound to d(CCAGGCCTGG). J Am Chem Soc 1997;119:9603–9613.
52. Caceres-Cortes J, Sugiyama H, Ikudome K, Saito I, Wang AH-J. Interactions of deglycosylated cobalt(III)- pepleomycin (green form) with DNA based on NMR structural studies. Biochemistry 1997;36:9995–10005.
53. Rajani C, Kincaid JR, Petering DH. The presence of two modes of binding to calf thymus DNA by metal- free bleomycin: A low frequency Raman study. Biopolymers 1999;52:129–146.
54. (a) Manderville RA, Ellena JF, Hecht SM. Solution structure of a Zn(II)*bleomycin A5-D(CGCTAGCG)2 complex. J Am Chem Soc 1994;116:10851–10852. (b) Manderville RA, Ellena JF, Hecht SM. Interactionof Zn(II)*bleomycin with d(CGCTAGCG)2—A binding model based on NMR experiments and restrainedmolecular dynamics calculations. J Am Chem 1995;117:7891–7903.
55. Abraham AT, Zhou X, Hecht SM. DNA cleavage by Fe(II)center dot bleomycin conjugated to a solid support. J Am Chem Soc 1999;121:1982–1983 & 4092.
56. Hoehn ST, Junker H-D, Bunt RC, Turner CJ, Stubbe J. Solution structure of Co(III)-bleomycin-OOH bound to a phosphoglycolate lesion containing oligonucleotide: Implications for bleomycin-induced double-strand DNA cleavage. Biochemistry 2001;40:5894–5905.
57. (a) Noggle JH, Schirmer RE. The nuclear overhauser effect. New York, NY: Academic; 1971. (b) Neuhaus D, Williamson MP. The nuclear Overhauser effect in structural and conformational analysis. New York,NY: VCH; 1989.
58. (a) Sugiura Y. Bleomycin–iron complexes—Electron-spin resonance study, ligand effect, and implication for action mechanism. J Am Chem Soc 1980;102:5208–5215. (b) Burger RM, Peisach J, Horwitz SB.
Activated bleomycin. A transient complex of drug, iron, and oxygen that degrades DNA. J Biol Chem1981;256:1636–1644. (c) Veselov A, Sun HJ, Sienkiewicz A, Taylor H, Burger RM, Scholes CP. Ironcoordination of activated bleomycin probed by Q-band and X-band ENDOR—Hyperfine coupling toactivated O-17 oxygen, N-14 and exchangeable 1H. J Am Chem Soc 1995;117:7508–7512.
59. Takahashi S, Sam JW, Peisach J, Rousseau DL. Structural characterization of iron-bleomycin by resonance Raman spectroscopy. J Am Chem Soc 1994;116:4408–4413.
60. Burger RM, Kent TA, Horwitz SB, Mu¨nck E, Peisach J. Mo¨ssbauer study of iron bleomycin and its activation intermediates. J Biol Chem 1983;258:1559–1564.
61. Neese F, Zaleski JM, Zaleski KL, Solomon EI. Electronic structure of activated bleomycin: Oxygen intermediates in heme versus non-heme iron. J Am Chem Soc 2000;122:11703–11724.
62. Pillai RP, Lenkinski RE, Sakai TT, Geckle JM, Krishna NR, Glickson JD. Proton NMR study of iron(II)- bleomycin: Assignment of resonances by saturation transfer experiments. Biochem Biophys Res Commun1980;96:341–349.
63. Lehmann TE, Ming L-J, Rosen ME, Que L, Jr. NMR studies of the paramagnetic complex Fe(II)- bleomycin. Biochemistry 1997;36:2807–2816.
64. (a) La Mar GN, Horrocks WD, Jr., Holm RH, editors. NMR of paramagnetic molecules. New York: Academic; 1973. (b) Bertini I, Luchinat C. NMR of paramagnetic molecules in biological systems. MenloPark, CA: Benjamin/Cumming; 1986. (c) Berliner LJ, Reuben J, editors. NMR of paramagnetic molecules.
New York: Plenum; 1993. (d) La Mar GN, editor. Nuclear magnetic resonance of paramagnetic molecules.
NATO-ASI, Dordrecht, Netherlands: Kluwer; 1995. (e) Ming L-J. Nuclear magnetic resonance ofparamagnetic metal centers in proteins and synthetic complexes. In: Que L, Jr., editor. Physical methods inbioinorganic chemistry, spectroscopy, and magnetism. Sausalito, CA: University Science Books; 2000.
65. Wasinger EC, Zaleski KL, Hedman B, Hodgson KO, Solomon EI. X-ray absorption spectroscopic investigation of Fe(II)-peplomycin and peplomycin derivatives: The effect of axial ligation on Fe-pyrimidine back-bonding. J Biol Inorg Chem 2002;7:157–164.
66. Lehmann TE, Serrano ML, Que L, Jr. Coordination chemistry of Co(II)-bleomycin: Its investigation through NMR and molecular dynamics. Biochemistry 2000;39:3886–3898.
67. (a) Guajardo RJ, Hudson SE, Brown SJ, Mascharak PK. [Fe(PMA)]nþ (n ¼ 1,2)—Good models of Fe- bleomycins and examples of mononuclear nonheme iron complexes with significant O2-activationcapabilities. J Am Chem Soc 1993;115:7971–7977. (b) Guajardo RJ, Tan JD, Mascharak PK. Thesecondary amine group of bleomycin is not involved in intramolecular hydrogen-binding in activatedbleomycin. Inorg Chem 1994;33:2838–2840. (c) Guajardo RJ, Chavez F, Farinas ET, Mascharak PK.
Structural features that control oxygen activation at the nonheme iron site in Fe(II)-bleomycin—An analogstudy. J Am Chem Soc 1995;117:3883–3884.
68. Ghirlanda G, Scrimin P, Tecilla P, Toffoletti A. Amphiphilic copper(II) complexes modeled after the metal- complexation subunit of bleomycin antibiotics. Langmuir 1998;14:1646–1655 and References citedtherein.
69. (a) Kurosaki H, Ishikawa Y, Hayashi K, Sumi M, Tanaka Y, Goto M, Inada K, Taniguchi I, Shionoya M, Matsuo H, Sugiyama M, Kimura E. Synthesis, spectroscopic, redox properties, and DNA cleavage activityof low-spin iron(III) complexes of bleomycin analogues. Inorg Chem Acta 1999;294:56–61. (b) KurosakiH, Hayashi K, Ishikawa Y, Goto M, Inada K, Taniguchi I, Shionoya M, Kimura E. New robust bleomycinanalogues: Synthesis, spectroscopy, and crystal structures of the copper(II) complexes. Inorg Chem1999;38:2824–2832 & 4372. (c) References cited above.
70. Sam JW, Tang XJ, Magliozzo RS, Peisach J. Electrospray mass spectrometry of iron bleomycin— Investigation of the reaction of Fe(II)-bleomycin with iodosylbenzene. J Am Chem Soc 1995;117:1012–1018.
71. Ghirlanda G, Scrimin P, Tecilla P, Toffoletti A. Amphiphilic copper(II) complexes modeled after the metal- complexation subunit of bleomycin antibiotics. Langmuir 1998;14:1646–1655.
72. Kurosaki H, Hayashi K, Ishikawa Y, Goto M, Inada K, Taniguchi I, Shionoya M, Kimura E. New robust bleomycin analogues: Synthesis, spectroscopy, and crystal structures of the copper(II) complexes. InorgChem 1999;38:2824–2832.
73. (a) Stachelhaus T, Schneider A, Marahiel MA. Engineered biosynthesis of peptide antibiotics. Biochem Pharmacol 1996;52:177–186. (b) Konz D, Marahiel MA. How do peptide synthetases generate structuraldiversity? Chem Biol 1999;6:R39–R48. (c) Moffitt MC, Neilan BA. The expansion of mechanistic andorganismic diversity associated with non-ribosomal peptides. FEMS Microbiol Lett 2000;191:159–167.
(d) Weber T, Marahiel MA. Exploring the domain structure of modular nonribosomal peptide synthetases.
Struct Fold Design 2001;9:R3–R9.
74. Hopwood DA. Genetic contributions to understanding polyketide synthases. Chem Rev 1997;97:2465– 2498. (b) Shen B. Biosynthesis of aromatic polyketides. Top Curr Chem 2000;209:1–51. (c) Dittmann E,Neilan BA, Borner T. Molecular biology of peptide and polyketide biosynthesis in cyanobacteria. ApplMicrobiol Biotech 2001;57:467–473.
75. (a) Some examples and references therein: (a) McDaniel R, Ebert-Khosla S, Hopwood DA, Khosla C.
Engineered biosynthesis of novel polyketides. Science 1993;262:1546–1550. (b) Shen B, Hutchinson CR.
Enzymatic synthesis of a bacterial polyketide from acetyl and malonyl coenzyme A. Science1993;262:1535–1540. (c) Marsden AF, Caffrey P, Aparicio JF, Loughran MS, Staunton J, Leadlay PF.
Stereospecific acyl transfers on the erythromycin-producing polyketide synthase. Science 1994;263:378–380. (d) Pieper R, Luo G, Cane DE, Khosla C. Cell-free synthesis of polyketides by recombinanterythromycin polyketide synthases. Nature 1995;378:263–266. (e) Pelzer S, Reichert W, Huppert M,Heckmann D, Wohlleben W. Cloning and analysis of a peptide synthetase gene of the balhimycin producerAmycolatopsis mediterranei DSM5908 and development of a gene disruption/replacement system.
J Biotech 1997;56:115–128. (f) Saito M, Hori K, Kurotsu T, Kanda M, Saito Y. Three conserved glycineresidues in valine activation of gramicidin S synthetase 2 from Bacillus brevis. J Biochem 1995;117:276–282. (g) Nishizawa T, Asayama M, Fujii K, Harada K, Shirai M. Genetic analysis of the peptide synthetasegenes for a cyclic heptapeptide microcystin in Microcystis spp. J Biochem 1999;126:520–529. (h) Pelzer S,Reichert W, Huppert M, Heckmann D, Wohlleben W. Cloning and analysis of a peptide synthetase gene ofthe balhimycin producer Amycolatopsis mediterranei DSM5908 and development of a gene disruption/replacement system. J Biotech 1997;56:115–128. (i) Pelzer S, Su¨ssmuth R, Heckmann D, Recktenwald J,Huber P, Jung G, Wohlleben W. Identification and analysis of the balhimycin biosynthetic gene cluster and its use for manipulating glycopeptide biosynthesis in Amycolatopsis mediterranei DSM5908. AntimicrobAgent Chemother 1999;43:1565–1573. (j) Schembri MA, Neilan BA, Saint CP. Identification of genesimplicated in toxin production in the cyanobacterium Cylindrospermopsis raciborskii. Environ Toxicol2001;16:413–421. (k) Christiansen G, Dittmann E, Via Ordorika L, Rippka R, Herdman M, Bo¨rner T.
Nonribosomal peptide synthetase genes occur in most cyanobacterial genera as evidenced by theirdistribution in axenic strains of the PCC. Arch Microbiol 2001;176:452–458. (l) Tillett D, Dittmann E,Erhard M, von Dohren H, Borner T, Neilan BA. Structural organization of microcystin biosynthesis inMicrocystis aeruginosa PCC7806: An integrated peptide-polyketide synthetase system. Chem Biol2000;7:753–764. (m) Kwon H-J, Smith WC, Scharon AJ, Hwang SH, Kurth MJ, Shen B. C-O bondformation by polyketide synthases. Science 2002;297:1327–1330.
76. (a) Du L, Sa´nchez C, Chen M, Edwards DJ, Shen B. The biosynthetic gene cluster for the antitumor drug bleomycin from Streptomyces verticillus ATCC15003 supporting functional interactions betweennonribosomal peptide synthetases and a polyketide synthase. Chem Biol 2000;7:623–642. (b) Shen B,Du L, Sa´nchez C, Edwards DJ, Chen M, Murrell JM. Cloning and characterization of the bleomycinbiosynthetic gene cluster from Streptomyces verticillus ATCC15003(1). J Nat Prod 2002;65:422–431.
77. Keating TA, Walsh CT. Initiation, elongation, and termination strategies in plyketide and polypeptide antibiotic biosynthesis. Curr Opin Chem Biol 1999;3:598–606.
78. (a) Conti E, Franks NP, Brick P. Crystal structure of firefly luciferase throws light on a superfamily of adenylate formatting enzymes. Structure 1996;4:287–298. (b) Conti E, Stachelhaus T, Marahiel MA, BrickP. Structural basis for the activation of phenylalanine in the non-ribosomal biosynthesis of gramicidin.
EMBO J 1997;16:4174–4183. (c) Weber T, Baumgartner R, Renner C, Marahiel MA, Holak TA. Solutionstructure of PCP, a prototype for the peptidyl carrier domains of modular peptide synthetases. Structure2000;8:407–418.
79. (a) Marahiel MA. Protein templates for the biosynthesis of peptide antibiotics. Chem Biol 1997;4:561– 567. (b) Mootz HD, Marahiel MA. Design and application of multimodular peptide synthetases. Curr OpinBiotech 1999;10:341–348. (c) Kohli RM, Walsh CT, Burkart MD. Biomimetic synthesis and optimizationof cyclic peptide antibiotics. Nature 2002;418:658–661.
80. (a) Grundy WE, Goldstein AW, Rickher CJ, Hanes ME, Warren HB, Jr., Sylvester JC. Aureolic acid, a new antibiotic. 1. Microbiological studies. Antibiot Chemother 1953;3:1215–1217. (b) Remers WA. Thechemistry of antitumor antibiotics, Vol. 1. Chapter 3; 1979.
81. For example: (a) Ransohoff J, Martin BF, Medrek TJ, Harris MN, Golomb FM, Wright JC. Preliminary clinical study of mithramycin (nsc-24559) in primary tumors of the central nervous system. CancerChemother Rep 1965;49:51–57. (b) Kennedy BJ, Griffen WO, Jr., Lober P. Specific effect of mithramycinon embryonal carcinoma of the testis. Cancer 1965;18:1631–1636. (c) Sewell IA, Ellis H. A trial ofmithramycin in the treatment of advanced malignant disease. Brit J Cancer 1966;20:256–263. (d)Mackenzie AR, Duruman N, Whitmore WF, Jr. Mithramycin in metastatic urogenital cancer. J Urol1967;98:116–119. (e) Baum M. A clinical trial of mithramycin in the treatment of advanced malignantdisease. Brit J Cancer 1968;22:176–183. (f) Mealey J, Jr., Chen TT, Pedlow E. Brain tumor chemotherapywith mithramycin and vincristine. Cancer 1970;26:360–367. (g) Kofman S, Perlia CP, Economou SG.
Mithramycin in the treatment of metastatic Ewing's sarcoma. Cancer 1973;31:889–893. (h) Stamp TC,Child JA, Walker PG. Treatment of osteolytic myelomatosis with mithramycin. Lancet 1975;1(7909):719–722. (i) Schold SC, Jr., Bigner DD. Treatment of five subcutaneous human glioma tumor lines in athymicmice with carmustine, procarbazine, and mithramycin. Cancer Treat Rep 1983;67:811–819. (j) Koller CA,Campbell VW, Polansky DA, Mulhern A, Miller DM. In vivo differentiation of blast-phase chronicgranulocytic leukemia. Expression of c-myc and c-abl protooncogenes. J Clin Invest 1985;76:365–369. (k)Johnson PR, Yin JA, Narayanan MN, Geary CG, Love EM, Cinkotai KI. Failure of mithramycin to controlthe myeloid blast phase of chronic granulocytic leukemia: A report on nine patients and review of theliterature. Hematol Oncol 1991;9:9–15.
82. For example: (a) Ryan WG. Mithramycin for Paget's disease of bone. New Engl J Med 1970;283:1171. (b) Frame B, Marel GM. Paget disease: A review of current knowledge. Radiology 1981;141:21–24. (c) ZajacAJ, Phillips PE. Paget's disease of bone: Clinical features and treatment. Clin Exp Rheumatol 1985;3:75–88. (d) Kanis JA, Gray RE. Long-term follow-up observations on treatment in Paget's disease of bone. ClinOrthop Rel Res 1987;217:99–125. (e) Smidt WR, Hadjipavlou AG, Lander P, Dzioba RB. An algorithmicapproach to the treatment of Paget's disease of the spine. Orthop Rev 1994;23:715–724.
83. For example: (a) Perlia CP, Gubisch NJ, Wolter J, Edelberg D, Dederick MM, Taylor SG III. Mithramycin treatment of hypercalcemia. Cancer 1970;25:389–394. (b) Singer FR, Neer RM, Murray TM, KeutmannHT, Deftos LJ, Potts JT, Jr. Mithramycin treatment of intractable hypercalcemia due to parathyroidcarcinoma. New Engl J Med 1970;283:634–636. (c) Harrington G, Olson KB, Horton J, Cunningham T,Wright A. Neoplasia, hypercalcemia, and mithramycin. New Engl J Med 1970;283:1172. (d) Ellas EG, Reynoso G, Mittelman A. Control of hypercalcemia with mithramycin. Ann Surgery 1972;175:431–435.
(e) Kiang DT, Loken MK, Kennedy BJ. Mechanism of the hypocalcemic effect of mithramycin. J ClinEndocrinol Metab 1979;48:341–344. (f) Ralston SH, Gardner MD, Dryburgh FJ, Jenkins AS, Cowan RA,Boyle IT. Comparison of aminohydroxypropylidene diphosphonate, mithramycin, and corticosteroids/calcitonin in treatment of cancer-associated hypercalcaemia. Lancet 1985;2(8461):907–910. (g) RalstonSH, Gallacher SJ, Dryburgh FJ, Cowan RA, Boyle IT. Treatment of severe hypercalcaemia withmithramycin and aminohydroxypropylidene bisphosphonate. Lancet 1988;2(8605):277. (h) Davis KD,Attie MF. Management of severe hypercalcemia. Crit Care Clin 1991;7:175–190. (i) Ostenstad B,Andersen OK. Disodium pamidronate versus mithramycin in the management of tumour-associatedhypercalcemia. Acta Oncol 1992;31:861–864.
84. (a) Lo Cascio V, Rossini M, Bertoldo F. The medical treatment of primary hyperparathyroidism. Rec Prog Med 1993;84:287–295. (b) Kinirons MT. Newer agents for the treatment of malignant hypercalcemia. AmJ Med Sci 1993;305:403–406. (c) Hadjipavlou AG, Gaitanis LN, Katonis PG, Lander P. Paget's disease ofthe spine and its management. Eur Spine J 2001;10:370–384.
85. Chatterjee S, Zaman K, Ryu H, Conforto A, Ratan RR. Sequence-selective DNA binding drugs mithramycin A and chromomycin A3 are potent inhibitors of neuronal apoptosis induced by oxidative stressand DNA damage in cortical neurons. Ann Neurol 2001;49:345–354.
86. (a) Berlin YA, Esipov SE, Olivomycin IV. The structure of olivomycin. Tetrahedron Lett 1966;14:1431– 1436. (b) Berlin YA, Esipov SE, Kolosov MN, Shemyakin MM. The structure of the olivomycin-chromomycin antibiotics. Tetrahedron Lett 1966;15:1643–1647.
87. Bakhaeva GP, Berlin YA, Boldyreva EF, Chuprunova OA, Kolosov MN, Soifer VS, Vasiljeva TE, Yartseva IV. The structure of aureolic acid (mithramycin). Tetra Lett 1968;32:3595–3598.
88. Thiem J, Meyer B. Studies on the structure of chromomycin A3 by 1H and 13C nuclear magnetic resonance spectroscopy. J Chem Soc Perkin II 1979;1331–1336. (b) Kam M, Shafer RH, Berman E. Solutionconformation of the antitumor antibiotic chromomycin A3 determined by two-dimensional NMRspectroscopy. Biochemistry 1988;27:3581–3588.
89. (a) Thiem J, Meyer B. Studies on the structure of olivomycin a and mithramycin by 1H and 13C nuclear magnetic resonance spectroscopy. Tetrahedron 1981;37:551–558. (b) Thiem J, Schneider G. Synthesis andstructure of the disaccharide fragment B-A of mithramycin. Angew Chem Int Edit 1983;22:58–59. (c)Thiem J, Schneider G, Sinnwell V. Synthesis of olivosyl-oliosides and spectroscopic structure assignmentof mithramycin. Liebigs Ann Chem 1986;814–824. (d) Thiem J, Scho¨ttmer B. Beta-glycosylation in 2-deoxysaccharides—Convergent synthesis of the oligosaccharides of mithramycin. Angew Chem Int Edit1987;26:555–557.
90. Wohlert SE, Kunzel E, Machinek R, Mendez C, Salas JA, Rohr J. The structure of mithramycin reinves- tigated. J Nat Prod 1999;62:119–121.
91. (a) Itzhaki L, Weinberger S, Livnah N, Berman E. A unique binding cavity for divalent cations in the DNA– metal–chromomycin A3 complex. Biopolymers 1990;29:481–489. (b) Demicheli L, Albertini J-P,Garnier-Suillerot A. Interaction of mithramycin with DNA. Evidence that mithramycin binds to DNA as adimer in a right-handed screw conformation. Eur J Biochem 1991;198:333–338.
92. (a) Kersten W, Kersten H, Szybalsky W. Physicochemical properties of complexes between deoxyribonucleic acid and antibiotics which affect ribonucleic acid synthesis (actinomycin, daunomycin,cinerubin, nogalamycin, chormomycin, mithramycin, and olivomycin). Biochemistry 1966;5:236–244.
(b) Nayak L, Sirsi M, Podder SK. Role of magnesium ion on the interaction between chromomycin A3 anddeoxyribonucleic acid. FEBS Lett 1973;30:157–162. (c) Goldberg IH, Friedmann PA. Antibiotics andnucleic acids. Annu Rev Biochem 1971;40:775–810. (d) Demicheli L, Garnier-Suillerot A. Mithramycincannot bind to left-handed poly(dG-m5dC) in the presence of Mg2þ ion. Biochem Biophys Res Commun1991;177:511–517. (e) Demicheli L, Albertini JP, Garnier-Suillerot A. Interaction of mithramycin withDNA. Evidence that mithramycin binds to DNA as a dimer in a right-handed screw conformation. Eur JBiochem 1991;198:333–338. (f) Huang HW, Li D, Cowan JA. Biostructural chemistry of magnesium.
Regulation of mithramycin–DNA interactions by Mg2þ coordination. Biochimie 1995;77:729–738.
93. (a) Aich P, Dasgupta D. Role of Mg þþ in the mithramycin–DNA interaction: Evidence for two types of mithramycin–Mg þþ complex Biochem. Biophys Res Commun 1990;173:689–696. (b) Aich P, Shen R,Dasgupta D. Role of magnesium ion in the interaction between chromomycin A3 and DNA: Binding ofchromomycin A3–Mg2þ complexes with DNA. Biochemistry 1992;31:2988–2997. (c) Aich P, Sen R,Dasgupta D. Interaction between antitumor antibiotic chromomycin A3 and Mg2þ. I. Evidence for theformation of two types of chromomycin A3–Mg2þ complexes. Chem-Biol Inter 1992;83:23–33. (d) AichP, Dasgupta D. Role of magnesium ion in mithramycin–DNA interaction: Binding of mithramycin–Mg2þcomplexes with DNA. Biochemistry 1995;34:1376–1385. (e) Majee S, Sen R, Guha S, Bhattacharyya D,Dasgupta D. Differential interactions of the Mg2þ complexes of chromomycin A3 and mithramycin with poly(dG-dC)  poly(dC-dG) and poly(dG)  poly(dC). Biochemistry 1997;36:2291–2299. (f) ChakrabartiS, Mir MA, Dasgupta D. Differential interactions of antitumor antibiotics chromomycin A2 andmithramycin with d(TATGCATA)2 in presence of Mg2þ . Biopolymers 2001;62:131–140.
94. (a) Berman E, Brown SC, James TL, Shafer RH. NMR studies of chromomycin A3 interaction with DNA.
Biochemistry 1985;24:6887–6893. (b) Gao X, Patel DJ. Solution structure of the chromomycin–DNAcomplex. Biochemistry 1989;28:751–762. (c) Gao XL, Patel DJ. Antitumor drug–DNA interactions:NMR studies of echinomycin and chromomycin complexes. Quart Rev Biophys 1989;22:93–138. (d) GaoX, Patel DJ. Chromomycin dimer–DNA oligomer complexes. Sequence selectivity and divalent cationspecificity. Biochemistry 1990;29:10940–10956. (e) Banville DL, Keniry MA, Kam M, Shafer RH. NMRstudies of the interaction of chromomycin A3 with small DNA duplexes. Binding to GC-containingsequences. Biochemistry 1990;29:6521–6534. (f) Banville DL, Keniry MA, Shafer RH. NMR inves-tigation of mithramycin A binding to d(ATGCAT)2: A comparative study with chromomycin A3.
Biochemistry 1990;29:9294–9304. (g) Demicheli C, Albertini JP, Garnier-Suillerot A. Interaction ofmithramycin with DNA. Evidence that mithramycin binds to DNA as a dimer in a right-handed screwconformation. Eur J Biochem 1991;198:333–338. (h) Gao X, Mirau D, Patel DJ. Structure refinementof the chromomycin dimer–DNA oligomer complex in solution. J Mol Biol 1992;223:259–279. (i) SastryM, Patel DJ. Solution structure of the mithramycin dimer–DNA complex. Biochemistry 1993;32:6588–6604.
95. (a) van Dyke MW, Dervan PB. Chromomycin, mithramycin, and olivomycin binding sites on heterogeneous deoxyribonucleic acid. Footprinting with (methidiumpropyl-EDTA)iron(II). Biochemistry1983;22:2373–2377. (b) Fox KR, Howarth NR. Investigations into the sequence-selective binding ofmithramycin and related ligands to DNA. Nucleic Acids Res 1985;13:8695–8714. (c) Fox KR, Waring MJ.
Footprinting reveals that nogalamycin and actinomycin shuffle between DNA binding sites. Nucleic AcidsRes 1986;14:2001–2014. (d) Cons BMG, Fox KR. Interaction of mithramycin with metal ions and DNA.
Biochem Biophys Res Commun 1989;160:517–524. (e) Cons BMG, Fox KR. High resolution hydroxylradical footprinting of the binding of mithramycin and related antibiotics to DNA. Nucleic Acids Res1989;17:5447–5459. (f) Cons BMG, Fox KR. Effects of the antitumor antibiotic mithramycin on thestructure of repetitive DNA regions adjacent to its GC-rich binding site. Biochemistry 1991;30:6314–6321. (g) Stankus A, Goodisman J, Dabrowick JC. Quantitative footprinting analysis of the chromomycinA3–DNA interaction. Biochemistry 1992;31:9310–9318. (h) Bailey C, Waring MJ. Transferring thepurine 2-amino group from guanines to adenines in DNA changes the sequence-specific binding ofantibiotics. Nucleic Acids Res 1995;23:885–892.
96. (a) Miller DM, Polansky DA, Thomas SD, Ray R, Campbell VW, Sanchez J, Koller CA. Mithramycin selectively inhibits transcription of G-C containing DNA. Am J Med Sci 1987;294:388–394. (b) Ray R,Snyder RC, Thomas S, Koller CA, Miller DM. Mithramycin blocks protein binding and function of theSV40 early promoter. J Clin Invest 1989;83:2003–2007. (c) Nehls MC, Brenner DA, Gruss HJ, Dierbach H,Mertelsmann R, Herrmann F. Mithramycin selectively inhibits collagen-alpha 1(I) gene expression inhuman fibroblast. J Clin Invest 1993;92:2916–2921. (d) Ihn H, Tamaki K. Increased phosphorylation oftranscription factor Sp1 in scleroderma fibroblasts: Association with increased expression of the type Icollagen gene. Arthritis Rheum 2000;43:2240–2247.
97. (a) Baker VV, Shingleton HM, Hatch KD, Miller DM. Selective inhibition of c-myc expression by the ribonucleic acid synthesis inhibitor mithramycin. Am J Obstet Gynecol 1988;158:762–767. (b) Ray R,Thomas S, Miller DM. Mithramycin selectively inhibits the transcriptional activity of a transfected humanc-myc gene. Am J Med Sci 1990;300:203–208. (c) Snyder RC, Ray R, Blume S, Miller DM. Mithramycinblocks transcriptional initiation of the c-myc P1 and P2 promoters. Biochemistry 1991;30:4290–4297. (d)Campbell VW, Davin D, Thomas S, Jones D, Roesel J, Tran-Patterson R, Mayfield CA, Rodu B, Miller DM,Hiramoto RA. The G-C specific DNA binding drug, mithramycin, selectively inhibits transcription of thec-myc and c-ha-ras genes in regenerating liver. Am J Med Sci 1994;307:167–172. (e) Chen SJ, Chen YF,Miller DM, Li H, Oparil S. Mithramycin inhibits myointimal proliferation after balloon injury of the ratcarotid artery in vivo. Circulation 1994;90:2468–2473.
98. Hardenbol P, Van Dyke MW. In vitro inhibition of c-myc transcription by mithramycin. Biochem Biophys Res Commun 1992;185:553–558.
99. Jones DE, Jr., Cui DM, Miller DM. Expression of beta-galactosidase under the control of the human c-Myc promoter in transgenic mice is inhibited by mithramycin. Oncogene 1995;10:2323–2330.
100. (a) Keniry MA, Brown SC, Berman E, Shafer RH. NMR studies of the interaction of chromomycin A3 with small DNA duplexes I. Biochemistry 1987;26:1058–1067. (b) Shafer RH. High resolution NMR studies ofchromomycin–oligonucleotide interactions: Structure and sequence specificity. Biochem Pharm1988;37:1783–1784. (c) Shafer RH, Roques BP, LePecq JB, Delepierre M. Chromomycin A3 binds toleft-handed poly(dG-m5dC). Eur J Biochem 1988;173:377–382. (d) Sastry M, Fiala R, Patel DJ. Solution structure of mithramycin dimers bound to partially overlapping sites on DNA. J Mol Biol 1995;251:674–689.
101. Demicheli C, Garnier-Suillerot A. Mithramycin: A very strong metal chelating agent. Biochim Biophys 102. Durlach J, Durlach J, Bara M, Guietbara A. Magnesium and its relationship to oncology. Met Ions Biol Syst 103. Sedov KA, Sorokina IB, Berlin YA, Kolosov MN. Olivomycin and related antibiotics. 23. Effect of olivomycins, chromomycins, and mithramycin on leukemia La of mice of strain C57B1. Antibiotiki(Moscow) 1969;14:721–725.
104. (a) Silva DJ, Kahne DE. Studies of the 2:1 chromomycin A3–Mg2þ complex in methanol: Role of the carbohydrates in complex formation. J Am Chem Soc 1993;115:7962–7970. (b) Silva DJ, Goodnow R, Jr.,Kahne D. The sugars in chromomycin A3 stabilize the Mg2þ–dimer complex. Biochemistry 1993;32: 463–471.
105. Keniry MA, Banville DL, Simmonds PM, Shafer R. Nuclear magnetic resonance comparison of the binding sites of mithramycin and chromomycin on the self-complementary oligonucleotided(ACCCGGGT)2. Evidence that the saccharide chains have a role in sequence specificity. J Mol Biol1993;231:753 – 767.
106. Silva DJ, Kahne D. Chromomycin A3 as a blueprint for designed metal complexes. J Am Chem Soc 107. Roush WR, Lin X-F. Studies on the synthesis of aureolic acid antibiotics—Highly stereoselective synthesis of aryl 2-deoxy-beta-glycosides via the mitsunobu reaction and synthesis of the olivomycin A-Bdisaccharide. J Am Chem Soc 1995;117:2236–2250, and references therein.
108. Sastry M, Fiala R, Patel DJ. Solution structure of mithramycin dimers bound to partially overlapping sites on DNA. J Mol Biol 1995;251:674–689.
109. (a) Gochin M. A high-resolution structure of a DNA–chromomycin–Co(II) complex determined from pseudocontact shifts in nuclear magnetic resonance. Structure Fold Des 2000;8:441–452. (b) Desvaux H,Gochin M. Coherence transfer between nuclear spins in paramagnetic systems: Effects of nucleus-electrondipole-dipole cross-correlation. Mol Phys 1999;96:1317–1333. (c) Gochin M. Nuclear magneticresonance characterization of a paramagnetic DNA–drug complex with high spin cobalt; assignment ofthe 1H and 31P NMR spectra, and determination of electronic, spectroscopic, and molecular properties.
J Biomol NMR 1998;12:243–257. (d) Gochin M. Nuclear magnetic resonance studies of a paramagneticmetallo DNA complex. J Am Chem Soc 1997;119:3377–3378. (e) Tu K, Gochin M. Structure deter-mination by restrained molecular dynamics using NMR pseudocontact shifts as experimentally determinedconstraints. J Am Chem Soc 1999;121:9276–9285.
110. Marsh WS, Wesel EM, Garretson AL. Streptonigrin, an antitumor agent produced by strains of Streptomyces flocculus. 1. Microbiological studies. Antibiot Chemother 1961;11:151.
111. Several early in vivo and in vitro studies were published in Antibiot Chemother 1961;11(3):147–189.
112. (a) McBride TJ, Oleson JJ, Woolf D. The activity of streptonigrin against the Rauscher murine leukemia virus in vivo. Cancer Res 1966;26A:727–732. (b) White HL, White JR. Lethal action and metabolic effectsof streptonigrin on Escherichia coli. Mol Pharmacol 1968;4:549–565. (c) Livingston RB, Carter SK.
Single agents in cancer chemotherapy. New York: Plenum; 1970. pp 389–392. (d) Chirigos MA, PearsonJW, Papas TS, Woods WA, Wood HBJ, Spahn G. Effect of streptonigrin (NSC-45383) and analogs ononcornavirus replication and DNA polymerase activity. Cancer Chemother Rep 1973;57:305–309. (e)Woods WA, Massicot JG, Webb JH, Chirigos MA. Inhibitory effect of streptonigrin on a murine sarcomavirus-induced tumor cell line (MSC) and selection of drug-resistant clones. In Vitro 1973;9:24–30. (f)Brazhnikova MG, Dudnik YV. Methods of development of new anticancer drugs National Cancer InstituteMonograph: USA-USSR; 1975. pp 207–212.
113. Bolza´n AD, Bianchi NO, Bianchi MS. Chromosomal response of insect and mammalian cells to streptonigrin: A comparative study. Environ Mol Mutagen 1998;32:331–335.
114. Oernin D, Bay JO, Uhrhammer N, Bignon YJ. AT heterozygote cells exhibit intermediate levels of apoptosis in response to streptonigrin and etoposide. Eur J Cancer 1999;35:1030–1035.
115. (a) Lown JW, Sim S-K. Studies related to antitumor antibiotics. 7. Synthesis of streptonigrin analogs and their single-strained scission of DNA. Can J Chem 1976;54:2563–2572. (b) Lown JW, Sim SK. Studiesrelated to antitumor antibiotics. Part VIII. Cleavage of DNA by streptonigrin analogues and the relationshipto antineoplastic activity. Can J Biochem 1976;54:446–452. (c) Rao KV, Beach JW. Streptonigrin andrelated compounds. 5. Synthesis and evaluation of some isoquinoline analogues. J Med Chem1991;34:1871–1879. (d) Boger DL, Cassidy KC, Nakahara S. Total synthesis of streptonigrone. J AmChem Soc 1993;115:10733–10741. (e) Holzapfel CW, Dwyer C. o-nitrophenyltriflates in quinolinesynthesis: Easy access to a streptonigrin synthon. Heterocycles 1998;48:215–219. (f) Crous R, Dwyer C, Holzapfel CW. Cross coupling strategies toward the synthesis of the streptonigrin CD moiety. Heterocycles1999;51:721–726. (g) Kimber M, Anderberg PI, Harding MM. Synthesis of ABC analogues of theantitumor antibiotic streptonigrin. Tetrahedron 2000;56:3575–3581.
116. Hajdu J. Interaction of metal ions with streptonigrin and biological properties of the complexes. Metal Ions Biol Syst 1985;19:53–81.
117. Harding MM, Long GV. Interaction of the antitumor antibiotic streptonigrin with metal ions and DNA. Curr Med Chem 1997;4:405–420.
118. (a) White JR. Streptonigrin-transition metal complexes: Binding to DNA and biological activity. Biochem Biophys Res Commun 1977;77:387–391. (b) Rao KV. Interaction of streptonigrin with metals and withDNA. J Pharm Sci 1979;68:853–856. (c) Marcazzan M, Gatto B, Sissi C, Zagotto G, Capranico G, PalumboM. Further insight into the Zn2þ -mediated binding of streptonigrin to DNA. Farmaco 1998;53:645–649.
119. Bolza´n AD, Bianchi MS. Genotoxicity of streptonigrin: A review. Mutation Res 2001;488:25–37.
120. (a) Hajdu J, Armstrong EC. Interaction of metal ions with streptonigrin. 1. Formation of copper(II) and zinc(II) complexes of the anti-tumor antibiotic. J Am Chem Soc 1981;103:232–234. (b) Moustatih A,Garnier-Suillerot A. Bifunctional antitumor compounds: Synthesis and characterization of a Au(III)–streptonigrin complex with thiol-modulating properties. J Med Chem 1989;32:1426–1431. (c) FialloMML, Garnier-Suillerot A. Interaction of the antitumor drug streptonigrin with palladium(II) ions evidenceof the formation of a superoxo-palladium(II) –streptonigrin complex. Inorg Chem 1990;29:893–897.
121. Gutteridge J. Streptonigrin-induced deoxyribose degradation: Inhibition by superoxide dismutase, hydroxyl radical scavengers and iron chelators. Biochem Pharmacol 1984;33:3059–3062.
122. (a) White JR, Yeowell HN. Iron enhances the bactericidal action of streptonigrin. Biochem Biophys Res Commun 1982;106:407–411. (b) Merryfield ML, Lardy HA. Inhibition of phosphoenolpyruvatecarboxykinase by streptonigrin. Biochem Pharmacol 1982;31:1123–1129. (c) Yeowell HN, White JR.
Iron requirement in the bactericidal mechanism of streptonigrin. Antimicrob Agents Chemother1982;22:961–968. (d) Yeowell HN, White JR. Changes in streptonigrin lethality during adaptation ofEscherichia coli to picolinic acid. Correlation with intracellular picolinate and iron uptake. BiochimBiophys Acta 1984;797:302–311. (e) Williams PH, Carbonetti NH. Iron, siderophores, and the pursuit ofvirulence: Independence of the aerobactin and enterochelin iron uptake systems in Escherichia coli. InfectImmun 1986;51:942–947.
123. Sinha BK. Irreversible binding of reductively activated streptonigrin to nucleic acids in the presence of metals. Chem Biol Inter 1981;36:179– 188. (b) Sugiura Y, Kuwahara J, Suzuki T. DNA interac-tion and nucleotide sequence cleavage of copper-streptonigrin. Biochim Biophys Acta 1984;782:254 –261.
124. Chiu Y, Lipscomb WN. Molecular and crystal structure of streptonigrin. J Am Chem Soc 1975;97:2525– 125. Lown JW. The mechanism of action of quinone antibiotics. Mol Cell Biochem 1983;55:17–40.
126. (a) White HL, White JR. Interaction of streptonigrin with DNA in vitro. Biochim Biophys Acta 1966;123:648–651. (b) Cone R, Hasan SK, Lown JW, Morgan AR. The mechanism of the degradation ofDNA by streptonigrin. Can J Biochem 1976;54:219–223. (c) Bachur NR, Gordon SL, Gee MV. Generalmechanism for microsomal activation of quinone anti-cancer agents to free radicals. Cancer Res1978;38:1745–1750. (d) Cohen MS, Chai Y, Britigan B, Mckenna W, Adams J, Svendsen T, Bean K,Hassett D, Sparling F. Role of extracellular iron in the action of the quinone antibiotic streptonigrin:Mechanisms of killing and resistance of Neisseria gonorrhoeae. Antimicrob Agents Chemother1987;31:1507–1513. (e) Hassett DJ, Britigan BE, Svendsen T, Rosen GM, Cohen MS. Bacteria formintracellular free radicals in response to paraquat and streptonigrin. Demonstration of the potency ofhydroxyl radical. J Biol Chem 1987;262:13404–13408.
127. (a) Soedjak HS, Cano RE, Tran L, Bales BL, Hajdu J. Preparation and ESR spectroscopic characterization of the zinc(II) and cadmium(II) complexes of streptonigrin semiquinone. Biochim Biophys Acta1997;1335:305–314. (b) Soedjak HS, Bales BL, Hajdu J. Electron spin resonance of the semiquinone of theantitumor antibiotic streptonigrin and its metal complexes. Basic Life Sci 1988;49:203–210. (c) SoedjakHS, Bales BL, Hajdu J. Electron spin resonance of the semiquinone of the antitumor antibiotic streptonigrinand its metal complexes. In: Simic MG, Taylor KA, Sonntag CV, editors. Oxygen radicals in biology andmedicine. New York: Plenum; 1987.
128. (a) DeGraff W, Hahn SM, Mitchell JB, Krishna MC. Free radical modes of cytotoxicity of adriamycin and streptonigrin. Biochem Pharmacol 1994;48:1427–1435. (b) Krishna MC, Halevy RF, Zhang R, GutierrezA, Samuni A. Modulation of streptonigrin cytotoxicity by nitroxide SOD mimics. Free Rad Biol1994;17:379–388.
129. Long GV, Harding MM. A proton nuclear magnetic resonance study of the interaction of zinc(II) with the antitumor drug streptonigrin. J Chem Soc Dalton Trans 1996;549–552.
130. Long GV, Harding MM, Turner P. X-ray structure of the zinc complex of the central metal chelation site of the antitumor drug streptonigrin. Polyhedron 2000;19:1067–1071.
131. Long GV, Harding MM, Fan JY, Denny WA. Interaction of the antitumor antibiotic streptonigrin with DNA and oligonucleotides. Anti-Cancer Drug Design 1997;12:453–472.
132. Wei X, Ming L-J. NMR studies of metal complexes and DNA binding of the quinone-containing antibiotic streptonigrin. J Chem Soc Dalton Trans 1998;2793–2798.
133. (a) Arcamone F. Doxorubicin anticancer antibiotics. New York: Academic; 1981. (b) Weiss RB, Sarosy G, Clagett-Carr K, Russo M, Leyland-Jones B. Anthracycline analogs: The past, present, and future. CancerChemother Pharmacol 1986;18:185–197. (c) Lown JW, editor. Anthracycline and anthracenedione-basedanticancer agents. Amsterdam: Elsevier; 1988. (d) Lown JW. Discovery and development of anthracyclineantitumor antibiotics. Chem Soc Rev 1993;22:165–176. (e) Priebe W, editor. Anthracycline antibiotics:New analogues, methods of delivery, and mechanisms of action. ACS symposium series 547. WashingtonDC: American Chemical Society; 1995.
134. Richardson DS, Johnson SA. Anthracyclines in haematology: Preclinical studies, toxicity and delivery systems. Blood Rev 1997;11:201–223.
135. Kim BS, Moon SS, Hwang BK. Structure elucidation and antifungal activity of an anthracycline antibiotic, daunomycin, isolated from Actinomadura roseola. J Agricul Food Chem 2000;48:1875–1881.
136. (a) Steinherz LJ, Graham T, Hurwitz R, Sondheimer HM, Schwartz RG, Shaffer EM, Sandor G, Benson L, Williams R. Guidelines for cardiac monitoring of children during and after anthracycline therapy: Report ofthe Cardiology Committee of the Children's Cancer Study Group. Pediatrics 1992;89:942–949. (b) SingalPK, Iliskovic N, Li TM, Kumar D. Adriamycin cardiomyopathy: Pathophysiology and prevention. FASEBJ 1997;11:931–936.
137. Calabresi P, Parks RE, Jr. Chemotherapy of neoplastic diseases. In: Gilman AG, Goodman LS, Gilman A, editors. The pharmacological basis of therapeutics. New York: Macmillan; 1980. pp 1249–1313.
138. Jung K, Reszka R. Mitochondria as subcellular targets for clinically useful anthracyclines. Adv Drug Deliv 139. There are numerous examples of different anthracycline antibiotics that can be found in the literature with a simple key-word search for ‘‘anthracycline AND isolation.'' Some recent examples are: (a) Zagotto G,Gatto B, Moro S, Sissi C, Palumbo M. Anthracyclines: Recent developments in their separation andquantitation. J Chromat B Biomed Sci Appl 2001;764:161–171. (b) Kunnari TJ, Ylihonko KP, Klika KD,Ma¨ntsa¨la¨ PI, Hakala JM. Hybrid anthracyclines from a genetically engineered Streptomyces galilaeusmutant. J Org Chem 2000;65:2851–2855. (c) Pfefferle CM, Breinholt J, Olsen CE, Kroppenstedt RM,Wellington EM, Gu¨rtler H, Fiedler HP. Kyanomycin, a complex of unusual anthracycline-phospholipidhybrids from Nonomuria species. J Nat Prod 2000;63:295–298. (d) Speitling M, Nattewan P, Yazawa K,Mikami Y, Gru¨n-Wollny I, Ritzau M, Laatsch H, Gra¨fe U. Demethyl mutactimycins, new anthracyclineantibiotics from Nocardia and Streptomyces strains. J Antibiot 1998;51:693–698. (e) Momose I, KinoshitaN, Sawa R, Naganawa H, Iinuma H, Hamada M, Takeuchi T. Nothramicin, a new anthracycline antibioticfrom Nocardia sp. MJ896-43F17. J Antibiot 1998;51:130–135. (f) Komatsu Y, Takahashi O, Hayashi H.
Identification of the anthracycline antibiotic 4-O-(beta-D-glucopyranuronosyl)-epsilon-rhodomycinone,produced by Streptomyces ruber JCM3131, as an up-regulator of MHC class-I molecules in B16/BL6 cells.
J Antibiot 1998;51:85–88. (g) Johdo O, Yoshioka T, Takeuchi T, Yoshimoto A. Isolation of newanthracyclines 10-O-rhodosaminyl beta-rhodomycinone and beta-isorhodomycinone from mild-acidtreated culture of obelmycin-producing Streptomyces violaceus. J Antibiot 1997;50:522–525. (h)Kawauchi T, Sasaki T, Yoshida K, Matsumoto H, Chen RX, Huang MY, Tsuchiya KS, Otani T. A newanthracycline antibiotic, IT-62-B, converts the morphology of ras-transformed cells back to normal:Taxonomy, fermentation, isolation, structure elucidation, and biological characterization. J Antibiot1997;50:297–303. (i) Kim HS, Kim YH, Yoo OJ, Lee JJ. Aclacinomycin X, a novel anthracyclineantibiotic produced by Streptomyces galilaeus ATCC 31133. Biosci Biotech Biochem 1996;60:906–908.
(j) Ishigami K, Hayakawa Y, Seto H. Cororubicin, a new anthracycline antibiotic generating active oxygenin tumor cells. J Antibiot 1994;47:1219–1225.
140. (a) Giannini G. Fluorinated anthracyclines: Synthesis and biological activity. Curr Med Chem 2002;9:687– 712. (b) Animati F, Arcamone F, Bigioni M, Capranico G, Caserini C, De Cesare M, Lombardi P, Pratesi G,Salvatore C, Supino R, Zunino F. Biochemical and pharmacological activity of novel 8-fluoroanthracy-clines: Influence of stereochemistry and conformation. Mol Pharmacol 1996;50:603–609. (c) Tonkin KC,Boston RC, Brownlee RT, Phillips DR. Fluorinated anthracyclines: Interactions with DNA. Invest NewDrugs 1990;8:355–363. (d) Tonkin KC, Brownlee RT, Zunino F, Phillips DR. Fluorinated anthracyclines:Synthesis and biological activity. Invest New Drugs 1990;8:1–6. (e) Chaires JB, Leng F, Przewloka T, FoktI, Ling YH, Perez-Soler R, Priebe W. Structure-based design of a new bisintercalating anthracyclineantibiotic. J Med Chem 1997;40:261–266.
141. (a) Courseille C, Busetta B, Geoffre S, Hospital M. Acta Cryst 1979;B35:764–767. (b) Arcamone F, Franceschi G, Orezzi P, Penco S, Mondelli R. The structure of daunomycin. Tetra Lett 1968;30:3349–3352. (c) Arcamone F, Cassinelli G, Franceschi G, Orezzi P, Mondelli R. The total absolute configuration ofdaunomycin. Tetra Lett 1968;30:3353–3356.
142. Powis G. Metabolism and reactions of quinoid anticancer agents. In: Powis G, editor. Anticancer drugs: Reactive metabolism and drug interactions. Oxford: Pergamon; 1994. pp 273–386.
143. Gewirtza DA. A critical evaluation of the mechanisms of action proposed for the antitumor effects of the anthracycline antibiotics adriamycin and daunorubicin. Biochem Pharmacol 1999;57:727–741.
144. Some recently determined structures: (a) Moore MH, Hunter WN, d'Estaintot BL, Kennard O. DNA–drug interactions. The crystal structure of d(CGATCG) complexed with daunomycin. J Mol Biol 1989;206:693–705. (b) Frederick CA, Williams LD, Ughetto G, van der Marel GA, van Boom JH, Rich A, Wang AH.
Structural comparison of anticancer drug–DNA complexes: Adriamycin and daunomycin. Biochemistry1990;29:2538–2549. (c) Williams LD, Frederick CA, Ughetto G, Rich A. Ternary interactions of sperminewith DNA:4 0-epiadriamycin and other DNA:anthracycline complexes. Nucleic Acids Res 1990;18:5533–5541. (d) Langlois d'Estaintot B, Gallois B, Brown T, Hunter WN. The molecular structure of a 4 0-epiadriamycin complex with d(TGATCA) at 1.7 A ˚ resolution: Comparison with the structure of 40- epiadriamycin d(TGTACA) and d(CGATCG) complexes. Nucleic Acids Res 1992;20:3561–3566. (e)Leonard GA, Hambley TW, Mcauleyhecht K, Brown T, Hunter WN. Anthracycline–DNA interactions atunfavorable base-pair triplet-binding sites—Structure of d(CGGCCG)-daunomycin and d(TGGCCA)-adriamycin complexes. Acta Crystallogr D Biol Crystallogr 1993;49:458–467. (f) Dautant A, Langloisd'Estaintot B, Gallois B, Brown T, Hunter WN. A trigonal form of the idarubicin:d(CGATCG) complex;crystal and molecular structure at 2.0 A ˚ resolution. Nucleic Acids Res 1995;23:1710–1716. (g) Smith CK, Davies GJ, Dodson EJ, Moore MH. DNA–nogalamycin interactions: The crystal structure of d(TGATCA)complexed with nogalamycin. Biochemistry 1995;34:415–425. (h) Smith CK, Brannigan JA, Moore MH.
Factors affecting DNA sequence selectivity of nogalamycin intercalation: The crystal structure ofd(TGTACA)2-nogalamycin2. J Mol Biol 1996;263:237–258. (i) Gao Y-G, Robinson H, Wijsman ER, vander Marel GA, van Boom JH, Wang AH-J. Binding of daunomycin to beta-D-glucosylated DNA found inprotozoa Trypanosoma brucei studied by X-ray crystallography. J Am Chem Soc 1997;119:1496–1497. (j)Podell ER, Harrington DJ, Taatjes DJ, Koch TH. Crystal structure of epidoxorubicin-formaldehyde virtualcrosslink of DNA and evidence for its formation in human breast-cancer cells. Acta Crystallogr D1999;55:1516–1523.
145. (a) Searle MS, Bicknell W. Interaction of the anthracycline antibiotic nogalamycin with the hexamer duplex d(5 0-GACGTC)2. An NMR and molecular modeling study. Eur J Biochem 1992;205:45–58. (b)Yang D, Wang AH. Structure by NMR of antitumor drugs aclacinomycin A and B complexed tod(CGTACG). Biochemistry 1994;33:6595–6604. (c) Caceres-Cortes J, Wang AH. Binding of theantitumor drug nogalamycin to bulged DNA structures. Biochemistry 1996;35:616–625. (d) Williams HE,Searle MS. Structure, dynamics, and hydration of the nogalamycin-d(ATGCAT)2 complex determined byNMR and molecular dynamics simulations in solution. J Mol Biol 1999;290:699–716. (e) Robinson H,Priebe W, Chaires JB, Wang AH. Binding of two novel bisdaunorubicins to DNA studied by NMRspectroscopy. Biochemistry 1997;36:8663–8670.
146. Chakrabarti S, Mahmood A, Kassis AI, Bump EA, Jones AG, Makrigiorgos GM. Generation of hydroxyl radicals by nucleohistone-bound metal–adriamycin complexes. Free Rad Res 1996;25:207–220.
147. Olson RD, Mushlin PS. Doxorubicin cardiotoxicity—Analysis of prevailing hypotheses. FESEB J 1990;4:3076–3086. (b) Singal PK, Iliskovic N, Li T, Kumar D. Adriamycin cardiomyopathy:Pathophysiology and prevention. FASEB J 1997;11:931–936. (c) Lee V, Randhawa AK, Singal PK.
Adriamycin-induced myocardial dysfunction in vitro is mediated by free radicals. Am J Physiol1991;261:H989–H995. (d) Singal PK, Panagia V. Direct effects of adriamycin on the rat heart sarcolemma.
Res Commun Chem Pathol Pharmacol 1984;43:67–77.
148. De Jong J, Husken BCP, Beekman B, van der Vijgh WJF, Bast A. Radical formation by metal-complexes of anthracyclines and their metabolites—Is there a relation with cardiotoxicity? Eur J Pharm Sci 1994;2:229–237.
149. Minotti G, Cairo G, Monti F. Role of iron in anthracycline cardiotoxicity: New tunes for an old song? FASEB J 1999;13:199–212. (b) Olson RD, Mushlin PS, Brenner DE, Fleischer S, Cusack BJ, Chang BK,Boucek RJ, Jr. Doxorubicin cardiotoxicity may be caused by its metabolite, doxorubicinol. Proc Nat AcadSci USA 1988;85:3585–3589. (c) Boucek RJ, Jr., Olson RD, Brenner DE, Ogunbunmi EM, Inui M,Fleischer S. The major metabolite of doxorubicin is a potent inhibitor of membrane-associated ion pumps.
A correlative study of cardiac muscle with isolated membrane fractions. J Biol Chem 1987;262:15851–15856.
150. Martin RB. Tetracyclines and daunomycin. Met Ions Biol Sys 1985;19:19–52.
151. (a) Papakyriakou A, Anagnostopoulou A, Garnier-Suillerot A, Katsaros N. Interaction of uranyl ions with daunorubicin and adriamycin. Eur J Inorg Chem 2002;1146–1154. (b) Balestrieri E, Bellugi L, Boicelli A,Giomini M, Giuliani AM, Giustini M, Marciani L, Sadler PJ. Interaction of tin(IV) with doxorubicin. JChem Soc Dalton Trans 1997;4099–4105. (c) Pereira E, Fiallo MML, Garniersuillerot A, Kiss T,Kozlowski H. Impact of aluminum ions on adriamycin-type ligands. J Chem Soc Dalton Trans 1993:455–459.
152. (a) Greenaway FT, Dabrowiak JC. The binding of copper ions to daunomycin and adriamycin. J Inorg Biochem 1982;16:91–107. (b) Tachibana M, Iwaizumi M, Tero-Kubota SJ. EPR studies of copper(II) andcobalt(II) complexes of adriamycin. J Inorg Biochem 1987;30:133–140. (c) Tachibana M, Iwaizumi M.
EPR and UV-visible spectroscopic studies of copper(II) and cobalt(II) complexes of hydroxyanthraqui-nones. J Inorg Biochem 1987;30:141–151.
153. (a) Matzanke BF, Bill E, Butzlaff C, Trautwein AX, Winkler H, Hermes C, Nolting H-F, Barbieri R, Russo U. Evidence for polynuclear aggregates of ferric daunomycin. A Mo¨ssbauer, EPR, X-ray absorptionspectroscopy and magnetic susceptibility study. Eur J Biochem 1992;207:747–755. (b) Massoud SS,Jordan RB. Kinetic and equilibrium studies of the complexation of aqueous iron(III) by daunomycin,quinizarin, and quinizarin-2-sulfonate. Inorg Chem 1991;30:4851–4856. (c) Gelvan D, Berg E, Saltman P,Samuni A. Time-dependent modifications of ferric-adriamycin. Biochem Pharmacol 1990;39:1289–1295.
154. (a) Allman T, Lenkinski RE. Adriamycin complexes of Pd(II) and Pt(II). J Inorg Biochem 1987;30:35–43.
(b) Pasini A. A doxorubicin–Pt(II) complex. Chemistry and antitumor activity. Inorg Chim Acta1987;137:123–124. (c) Fiallo MML, Garnier-Suillerot A. Metal anthracycline complexes as a new class ofanthracycline derivatives—Pd(II)-adriamycin and Pd(II)-daunomycin complexes—Physicochemicalcharacteristics and antitumor activity. Biochemistry 1986;25:924–930.
155. (a) Lenkinski RE, Sierke S. Vist MR. Lanthanide complexes of adriamycin. J Less-Comm Met 1983;94: 359–365. (b) Lenkinski RE, Sierke S. The thermodynamics of lanthanide ion binding to adriamycin.
J Inorg Biochem 1985;24:59–67.
156. Wei X, Ming L-J. Comprehensive 2D 1H NMR studies of paramagnetic lanthanide(III) complexes of anthracycline antitumor antibiotics. Inorg Chem 1998;37:2255–2262.
157. (a) Myers CE, Gianni L, Simone CB, Klecker R, Greene R. Oxidative destruction of erythrocyte ghost membranes catalyzed by the doxorubicin–iron complex. Biochemistry 1982;21:1707–1713. (b) MuindiJRF, Sinha BK, Gianni L, Myers CE. Hydroxyl radical production and DNA damage induced byanthracycline–iron complex. FEBS Lett 1984;172:226–230. (c) Eliot H, Gianni L, Myers C. Oxidativedestruction of DNA by the adriamycin iron complex. Biochemistry 1984;23:928–936. (c) Beraldo H,Garnier-Suillerot A, Tosi L, Lavelle F. Iron(III)–adriamycin and Iron(III)–daunorubicin complexes:Physicochemical characteristics, interaction with DNA, and antitumor activity. Biochemistry 1985;24:284–289. (d) Cullinane C, Phillips DR. Induction of stable transcriptional blockage sites by adriamycin—GPC specificity of apparent adriamycin DNA adducts and dependence on iron(III) ions. Biochemistry1990;29:5638–5646. (e) Akman SA, Doroshow JH, Burke TG, Dizdaroglu M. DNA-base modificationsinduced in isolated human chromatin by NADH dehydrogenase-catalyzed reduction of doxorubicin.
Biochemistry 1992;31:3500–3506.
158. (a) Phillips DR, Carlyle GA. The effect of physiological levels of divalent metal ions on the interaction of daunomycin with DNA: Evidence of a ternary daunomycin-Cu2þ–DNA complex. Biochem Pharmacol1981;30:2021–2024. (b) Spinelli M, Dabrowiak JC. Interaction of copper(II) ions with the daunomycincalf thymus deoxyribonucleic acid complex. Biochemistry 1982;21:5862–5870. (c) Reenaway FT,Dabrowiak JC. The binding of copper ions to daunomycin and adriamycin. J Inorg Biochem 1982;16:91–107. (d) Mariam YH, Glover GP. Degradation of DNA by metalloanthracyclines: Requirement for metalions. Biochem Biophys Res Commun 1986;136:1–7.
159. Rowley DA, Halliwell B. DNA damage by superoxide-generating systems in relation to the mechanism of action of the anti-tumour antibiotic adriamycin. Biochim Biophys Acta 1983;761:86–93.
160. Houee-Levin C, Gardes-Albert M, Rouscilles A, Ferradini C, Hickel B. Intramolecular semiquinone disproportionation in DNA—Pulse radiolysis study of the one-electron reduction of daunorubicinintercalated in DNA. Biochemistry 1991;30:8216–8222.
161. Kiraly R, Martin RB. Metal binding to daunorubicin and quinizarin. Inorg Chim Acta 1982;67:13–18.
162. Haj-Tajeb HB, Fiallo MML, Garniersuillerot A, Kiss T, Kozlowski H. Anthracycline anticancer drugs as effective ligands for terbium(III) ions. J Chem Soc Dalton Trans 1994;3689–3693.
163. Zweier JL, Levy A. Electron paramagnetic resonance and Mo¨ssbauer studies of metal chelation by adriamycin. Magn Reson Chem 1995;33:S114–S122.
164. Capolongo F, Giomini M. The interactions of Fe3þ ions with adriamycin studied by 57Fe Mo¨ssbauer and electronic spectroscopies. J Inorg Biochem 1997;65:115–122.
165. (a) Fiallo MML, Garnier-Suillerot A, Matzanke B, Kozlowski H. How Fe3þ binds anthracycline antitumor compounds. The myth and the reality of a chemical sphinx. J Inorg Biochem 1999;75:105–115. (b) FialloMML, Drechsel H, Garnier-Suillerot A, Matzanke BF, Kozlowski H. Solution structure of iron(III)–anthracycline complexes. J Med Chem 1999;42:2844–2851.
166. Kostoryz EL, Yourtee DM. Oxidative mutagenesis of doxorubicin–Fe(III) complex. Mutation Res Gene Toxicol Enviro Mutagen 2001;490:131–139.
167. (a) Ortega R. Intracellular distributions of the anthracycline 4 0-iodo-4 0-deoxy-doxorubicin and essential trace metals using nuclear microprobe analysis. Polycyc Arom Comp 2000;21:99–108. (b) Fiallo M,Laigle A, Garniersuillerot A, Amirand C, Baillini JP, Chinsky L, Duquesne M, Jolles B, Sureeau F, TurpinPY, Vigny P. Interactions of iron–anthracycline complexes with living cells—A microspectrofluorometricstudy. Biochim Biophys Acta 1993;1177:236–244.
168. (a) Ming L-J. Paramagnetic lanthanide(III) ions as NMR probes for biomolecular structure and function. In: La Mar GN, editor. Nuclear magnetic resonance of paramagnetic molecules. Dordrecht, Netherlands:NATO-ASI, Kluwer; 1995. (b) Nieboer E. The lanthanide ions as structural probes in biological and modelsystems. Struct Bond 1975;22:1–47. (c) Horrocks WdeW. Lanthanide ion probes of biomolecularstructure. Adv Inorg Biochem 1982;4:201–260.
169. (a) Bunzli J-CG, Choppin GR. Lanthanide probes in life, chemical, and earth science. Amsterdam: Elsevier; 1989. (b) Evans CH. Biochemistry of the Lanthanides. New York: Plenum; 1990. (c) WilliamsRJP. The chemistry of lanthanide ions in solution and in biological systems. Struct Bond 1982;50:79–119.
170. Mclennan IJ, Lenkinski RE. The binding of Yb(III) to adriamycin. A 1H NMR relaxation study. J Am Chem 171. Ming L-J, Wei X. An ytterbium(III) complex of daunomycin, a model metal complex of anthracycline antibiotics. Inorg Chem 1994;33:4617–4618.
172. Wei X. Two-dimensional NMR studies of paramagnetic metallo-biomolecules: Metal-antibiotic drug complexes and protein structure determination. Ph.D. Dissertation, University of South Florida, 1996.
173. Courseille C, Busetta B, Geoffre S, Hospital M. Complex daunomycin-butanol. Acta Cryst B 174. Davtyan TK, Avanesyan LA, Gukasyan VZ, Aleksanyan YT. Effect of adriamycin and its complexes with transition metals on induction of immune response of human lymphocytes in culture. Bull Exp Biol Med1999;128:710–712.
175. Davtyan TK, Gyulkhandanyan AV, Gambarov SS, Avanessian LA, Alexanyan YT. The effects of adriamycin and adriamycin complexes with transitional metals on Ca2þ-dependent Kþ channels of humanerythrocytes. Biochim Biophys Acta 1996;1297:182–190.
176. Bland KI, Palin WE, Fraunhofer JA, Morris RR, Adcock RA, Tobin GR II. Experimental and clinical observations of the effects of cytotoxic chemotherapeutic drugs on wound healing. Ann Surg 1984;199:782–790.
177. Devereux DF, Thibault L, Boretos J, Brennan MF. The quantitative and qualitative impairment of wound healing by adriamycin. Cancer 1979;43:932–938.
178. (a) Muszynska A, Palka J, Gorodkiewicz E. The mechanism of daunorubicin-induced inhibition of prolidase activity in human skin fibroblasts and its implication to impaired collagen biosynthesis. ExpToxicol Pathol 2000;52:149–155. (b) Muszynska A, Wolczynski S, Palka J. The mechanism foranthracycline-induced inhibition of collagen biosynthesis. Eur J Pharmacol 2001;411:17–25.
179. Hannun YA, Foglesong RJ. The adriamycin–iron(III) complex is a potent inhibitor of protein kinase-C.
J Biol Chem 1989;264:9960–9989.
180. Monti E, Perletti G, Piccinini F, Monzini F, Morazzoni F. Interaction of Cu(II) and Cu(II)-anthracycline complexes with protein kinase C. Spectromagnetic assessment of the inhibitory effect. Inorg Chim Acta1993;205:181–184.
181. Gallego J, Varani G. Targeting RNA with small-molecule drugs: Therapeutic promise and chemical challenge. Acc Chem Res 2001;34:836–843.
182. Fourmy D, Yoshizawa S, Puglisi J. Structural basis of aminoglycoside action. In: Schroeder R, Wallis MG, editors. RNA-binding antibiotics. Landes: Bioscience; 2001.
183. Lynch SR, Recht MI, Puglisi JD. Biochemical and nuclear magnetic resonance studies of aminoglycoside– RNA complexes. Meth Enzymol 2000;317:240–261.
184. (a) Davies J, Gorini L, Davis BD. Misreading of RNA code words induced by aminoglycoside antibiotics.
Mol Pharmacol 1965;1:93–106. (b) Davies J, Davis BD. Misreading of ribonucleic acid code words in-duced by aminoglycoside antibiotics. The effect of drug concentration. J Biol Chem 1968;243:3312–3316.
185. (a) Moazed D, Noller HF. Interaction of antibiotics with functional sites in 16S ribosomal RAN. Nature 1987;327:389–394. (b) Woodcock J, Moazed D, Cannon M, Davies J, Noller HF. Interaction of antibioticswith A- and P-site-specific bases in 16S ribosomal RNA. EMBO J 1991;10:3099–3103.
186. Mei H-Y, Cui M, Heldsinger A, Lemrow SM, Loo JA, Sannes-Lowery KA, Sharmeen L, Czarnik AW.
Inhibitors of protein-RNA complexation that target the RNA: Specific recognition of humanimmunodeficiency virus type 1 TAR RNA by small organic molecules. Biochemistry 1998;37:14204–14212.
187. Rogers J, Chang AH, von Ahsen U, Schroeder R, Davies J. Inhibition of the self-cleavage reaction of the human hepatitis delta virus ribozyme by antibiotics. J Mol Biol 1996;259:916–925.
188. Earnshaw DJ, Gait MJ. Aminoglycosides and cleavage of the hairpin ribozyme. In: Schroeder R, Wallis MG, editors. RNA-binding antibiotics. Landes: Bioscience; 2001.
189. (a) Recht MI, Fourmy D, Blanchard SC, Dahlquist KD, Puglisi JD. RNA sequence determinants for aminoglycoside binding to an A-site rRNA model oligonucleotide. J Mol Biol 1996;262:421–436. (b)Blanchard SC, Fourmy D, Eason RG, Puglisi JD. rRNA chemical groups required for aminoglycosidebinding. Biochemistry 1998;37:7716–7724.
190. (a) Fourmy D, Yoshizawa S, Puglisi JD. Paromomycin binding induces a local conformational change in the A-site of 16 S rRNA. J Mol Biol 1998;277:333–345. (b) Fourmy D, Recht MI, Blanchard SC, PuglisiJD. Structure of the A site of Escherichia coli 16S ribosomal RNA complexed with an aminoglycosideantibiotic. Science 1996;247:1367–1371. (c) Yoshizawa S, Fourmy D, Puglisi JD. Structural origins ofgentamicin antibiotic action. EMBO J 1998;17:6437–6448.
191. (a) Lynch SR, Puglisi JD. Structure of a eukaryotic decoding region A-site RNA. J Mol Biol 2001;306: 1023–1035. (b) Fourmy D, Recht MI, Puglisi JD. Binding of neomycin-class aminoglycoside antibiotics tothe A-site of 16 S rRNA. J Mol Biol 1998;277:347–362.
192. (a) Clemons WM, Jr., May JL, Wimberly BT, McCutcheon JP, Capel MS, Ramakrishnan V. Structure of a bacterial 30S ribosomal subunit at 5.5 A resolution. Nature 1999;400:833–840. (b) Wimberly BT,Brodersen DE, Clemons WM, Jr., Morgan-Warren RJ, Carter AP, Vonrhein C, Hartsch T, Ramakrishnan V.
Structure of the 30S ribosomal subunit. Nature 2000;407:327–339.
193. Carter AP, Clemons WM, Brodersen DE, Morgan-Warren RJ, Wimberly BT, Ramakrishnan V. Functional insights from the structure of the 30S ribosomal subunit and its interactions with antibiotics. Nature2000;407:340–348.
194. Vicens Q, Westhof E. Crystal structure of paromomycin docked into the eubacterial ribosomal decoding A site. Structure 2001;9:647–658.
195. Beauclerk AA, Cundliffe E. Sites of action of two ribosomal RBA methylases responsible of resistance to aminoglycosides. J Mol Biol 1987;193:661–671.
196. Priuska EM, Schacht J. Formation of free radicals by gentamicin and iron and evidence for an iron/ gentamicin complex. Biochem Pharmacol 1995;50:1749–1752.
197. Priuska EM, Clark-Baldwin K, Pecoraro VL, Schacht J. NMR studies of iron–gentamicin complexes and the implications for aminoglycoside toxicity. Inorg Chim Acta 1998;273:85–91.
198. Muranaka H, Suga M, Sato K, Nakagawa K, Akaike T, Okamoto T, Maeda H, Ando M. Superoxide scavenging activity of erythromycin–iron complex. Biochem Biophys Res Commun 1997;232:183–187.
199. Jezowska-Bojczuk M. Copper(II)-lincomycin: Complexation pattern and oxidative activity. J Inorg 200. Jezowska-Bojczuk M, Lesniak W. Coordination mode and reactivity of copper(II) complexes with kasugamycin. J Inorg Biochem 2001;85:99–105.
201. Jezowska-Bojczuk M, Bal W, Kozlowski H. Kanamycin revisited: A combined potentiometric and spectroscopic study of copper(II) binding to kanamycin B. Inorg Chim Acta 1998;275-276:541–545.
202. Jezowska-Bojczuk M, Karaczyn A, Kozlowski H. Copper(II) binding to tobramycin: Potentiometric and spectroscopic studies. Carbohy Res 1998;313:265–269.
203. Jezowska-Bojczuk M, Karaczyn A, Bal W. Copper(II) binding to geneticin, a gentamycin analog. J Inorg 204. (a) Jezowska-Bojczuk M, Lesniak W, Bal W, Kozlowski H, Gatner K, Jezierski A, Sobczak J, Mangani S, Meyer-Klaucke W. Molecular mechanism of hydrogen peroxide conversion and activation by Cu(II)-amikacin complexes. Chem Res Toxicol 2001;14:1353–1362. (b) Jezowska-Bojczuk M, Bal W. Co-ordination of copper(II) by amikacin. Complexation equilibria in solution and oxygen activation by theresulting complexes. J Chem Soc Dalton Trans 1998;153–159.
205. Jezowska-Bojczuk M, Bal W, Kasprzak KS. Copper(II) interactions with an experimental antiviral agent, I-deoxynojirimycin, and oxygen activation by resulting complexes. J Inorg Biochem 1996;63:231–239.
206. (a) Sreedhara A, Cowan JA. Catalytic hydrolysis of DNA by metal ions and complexes. J Biol Inorg Chem APR 2001;6(4):337–347 (b) Sreedhara A, Freed JD, Cowan JA. Efficient inorganic deoxyribonucleases.
Greater than 50-million-fold rate enhancement in enzyme-like DNA cleavage. J Am Chem Soc 2000;122:8814–8824. (c) Sreedhara A, Cowan JA. Efficient catalytic cleavage of DNA mediated bymetalloaminoglycosides. Chem Commun 1998;1737–1738.
207. (a) Sreedhara A, Patwardhan A, Cowan JA. Novel reagents for targeted cleavage of RNA sequences: Towards a new family of inorganic pharmaceuticals. Chem Commun 1999;1147–1148. (b) Sreedhara A,Cowan JA. Targeted site specific cleavage of HIV-1 viral Rev responsive element by copper amino-glycosides. J Biol Inorg Chem 2001;6:166–172.
208. (a) Smith JT. Awakening the slumbering potential of the 4-quinolone antibacterials. Pharm J 1984;233: 299–305. (b) Cornett JB, Wentland MP. Quinolone antibacterial agents. Ann Rep Med Chem 1986;21:139–148. (c) Fernandes PB, Chu DTW. Quinolones. Ann Rep Med Chem 1987;22:117–126. (d) ChuDTW, Fernandes MP. Structure–activity relationships of the fluoroquinolones. Antimicrob AgentsChemother 1989;33:131–135. (e) Chu DT. Recent progress in novel macrolides, quinolones, and 2-pyridones to overcome bacterial resistance. Med Res Rev 1999;19:497–520.
209. (a) Wolfson JS, Hooper DC, editors. Quinolone antimicrobial agents. Washington, DC: American Society for Microbiology; 1993. (b) Andriole VT, editor. The quinolones, 3rd edn. San Diego: Academic; 2000.
210. Chu DTW, Hallas R, Clement JJ, Alder L, McDonald E, Plattner JJ. Synthesis and antitumor activities of quinolone antineoplastic agents. Drugs Exp Clin Res 1992;18:275–282.
211. (a) News of the Week. Drugmakers form emergency task force. C&EN News Nov. 2001;5:12. (b) Enserink M. Researchers question obsession with Cipro. Science 2001;294:759–760.
212. (a) Casparian JM, Luchi M, Morrat R, Hinthorn D. Quinolones and tendon ruptures. S Med J 2000;93:488– 491. (b) Litt JZ. Fluoroquinolones. S Med J 2000;93:525–526. (c) Fried S. Bitter pills. New York: BantamBooks; 1998.
213. (a) Yukinori K, Kyuichi M, Hideo H. Interaction of quinolones with metal cations in aqueous solution.
Chem Pharm Bull (Tokyo) 1996;44:1425–1430. (b) Mendoza-Dı´az G, Perez-Alonso R, Moreno-EsparzaR. Stability constants of copper(II) mixed complexes with some 4-quinolone antibiotics and (N–N) donors.
J Inorg Biochem 1996;64:207–214. (c) Issopoulos PB. Spectrophotometric determination of trace amountsof iron(III) with norfloxacin as complexing reagent. Analyst 1989;114:627–630. (d) Lee DS, Han HJ,Kim K, Park WB, Cho JK, Kim JH. Dissociation and complexation of fluoroquinolone analogs. PharmBiomed Anal 1994;12:157–164. (e) Turel I, Bukovec N. Complex formation between some metals anda quinolone family member (ciprofloxacin). Polyhedron 1996;15:269–275. (f) Turel I. The interactions ofmetal ions with quinolone antibacterial agents. Coord Chem Rev, online publication, 19 February 2002.
214. (a) Turel I, Leban I, Bukovec N. Synthesis, characterization, and crystal structure of a copper(II) complex with quinolone family member (ciprofloxacin): Bis(1-cyclopropyl-6-fluoro-1,4-dihydro-4-oxo-7-piper-azin-1-yl quinoline-3-carboxylate) copper(II) chloride hexahydrate. J Inorg Biochem 1994;56:273–282.
(b) Ruiz M, Ortiz R, Perello´ L, Latorre J, Server-Carrio´ J. Potentiometric and spectroscopic studies oftransition-metal ion complexes with a quinolone derivative (cinoxacin). Crystal structure of new Cu(II) andNi(II) cinoxacin complexes. J Inorg Biochem 1997;65:87–96. (c) Macı´as B, Villa MV, Rubio I, Castin˜eirasA, Borra´s J. Complexes of Ni(II) and Cu(II) with ofloxacin crystal structure of a new Cu(II) ofloxacincomplex. J Inorg Biochem 2001;84:163–170.
215. (a) Ho¨ffken G, Borner K, Glatzel PD, Koeppe P, Lode H. Reduced enteral absorption of ciprofloxacin in the presence of antacids. Eur J Clin Microbiol Infect Dis 1985;4:345–345. (b) Turel I, Sonc A, Zupancic M,Sepcic K, Turk T. The synthesis and biological activity of some magnesium(II) complexes of quinolones.
Met Based Drugs 2000;7:101–104.
216. (a) Gao F, Yang P, Xie J, Wang H. Synthesis, characterization and antibacterial activity of novel Fe(III), Co(II), and Zn(II) complexes with norfloxacin. J Inorg Biochem 1995;60:61–67.
217. Chulvi C, Mun˜oz MC, Perello´ L, Ortiz R, Mun˜oz MC, Arriortua MI, Via J, Urtiaga K, Amigo´ JM, Ochando LE. Coordination behavior of cinoxacine: Synthesis and crystal structure of tris (cinoxacinate)cobaltate(II)of sodium hexahydrate (HCx ¼ 1-ethyl-4(1H)-oxo-(1,3)dioxolo-(4,5g) cinnoline-3-carboxylic acid). JInorg Biochem 1991;42:133–138.
218. (a) Turel I, Leban I, Klintschar G, Bukovec N, Zalar S. Synthesis, crystal structure, and characterization of two metal-quinolone compounds. J Inorg Biochem 1997;66:77–82. (b) Turel I, Golic L, Bukovec P, GubinaM. Antibacterial tests of bismuth(III)-quinolone (ciprofloxacin, cf) compounds against Helicobacter pyloriand some other bacteria. Crystal structure of (cfH2)2[Bi2Cl10]n˜H2O. J Inorg Biochem 1998;71:53–60.
219. Alvarez EJ, Vartanian VH, Brodbelt JS. Metal complexation reactions of quinolones antibiotics in a quadruple ion trap. Anal Chem 1997;69:1147–1155.
220. Robles J, Martı´n-Polo J, A ´ lvarez-Valtierra L, Hinojosa L, Mendoza-Dı´az G. A theoretical-experimental study on the structure and activity of certain quinolones and the interaction of their Cu(II)-complexes on aDNA model. Metal-Based Drugs 2000;7:301–311.
221. (a) Levine C, Hiasa H, Marians K. DNA gyrase and topoisomerase IV: Biochemical activities, physiological roles during chromosome replication, and drug sensitivities. Biochim Biophys Acta 1998;1400:29–43. (b) Anderson V, Gootz T, Osheroff N. Topoisomerase IV catalysis and the mechanism ofquinolones action. J Biol Chem 1998;274:17879–17885. (c) Khodursky A, Cozzarelli N. The mechanismof inhibition of topoisomerase IV by quinolones antibacterials. J Biol Chem 1998, 273:27668–27677. (d)Marians K, Hiasa H. Mechanism of quinolones action: A drug-induced structural perturbation of the DNAprecedes strand cleavage by topoisomerase IV. J Biol Chem 1997;272:9401–9409. (e) Earlier publicationscited therein.
222. Elsea SH, Westergaard M, Burden DA, Lomenick JP, Osheroff N. Quinolones share a common interaction domain on topoisomerase II with other DNA cleavage-enhancing antineoplastic drugs. Biochemistry1997;36:2919–2924.
223. (a) Shen LL, Pernet AG. Mechanism of inhibition of DNA gyrase by analogues of nalidixic acid: The target of the drugs is DNA. Proc Nat Acad Sci 1985;82:307–311. (b) Shen LL, Baranowski J, Pernet AG.
Mechanism of inhibition of DNA gyrase by quinolone antibacterials: Specificity and cooperativity of drugbinding to DNA. Biochemistry 1989;28:3879–3885. (c) Shen LL, Mitscher LA, Sharma PN, O'DonnellTJ, Chu DWT, Cooper CS, Rosen T, Pernet AG. Mechanism of inhibition of DNA gyrase by quinoloneantibacterials: A cooperative drug-DNA binding model. Biochemistry 1989;28:3886–3894. (d) Shen LL,Kohlbrenner WE, Weigl D, Baranowski J. Mechanism of quinolone inhibition of DNA gyrase. Appearanceof unique norfloxacin binding sites in enzyme–DNA complexes. J Biol Chem 1989;264:2973–2978.
224. Palu G, Valisena S, Ciarrocchi G, Gatto B, Palumbo M. Quinolone binding to DNA is mediated by magnesium ions. Proc Nat Acad Sci USA 1992;89:9671–9675.
225. Tornaletti S, Pedrini AM. Studies on the interaction of 4-quinolones with DNA by DNA unwinding experiments. Biochim Biophys Acta 1988;949:279–287.
226. Lecomte S, Moreau NJ, Chenon M-T. NMR investigation of pefloxacin-cation-DNA interactions: The essential role of Mg2þ . Int J Pharm 1998;164:57–65.
227. Sissi C, Andreolli M, Cecchetti V, Fravolini A, Gatto B, Palumbo M. Mg2þ-mediated binding of 6-substituted quinolones to DNA: Relevance to biological activity. Bioorg Med Chem 1998;6:1555–1561.
228. Fan J-Y, Sun D, Yu H, Kerwin SM, Hurley LH. Self-assembly of a quinobenzoxazine–Mg2þ complex on DNA: A new paradigm for the structure of a drug–DNA complex and implications for the structure of thequinolone bacterial gyrase–DNA complex. J Med Chem 1995;38:408–424.
229. Polster J, Lachmann H. Spectrometric titrations. New York: VCH; 1989.
230. Kwok Y, Zeng Q, Hurley LH. Structural insight into a quinolone–topoisomerase II–DNA complex.
Further evidence for a 2:2 quinobenzoxazine-Mg2þ self-assembly model formed in the presence oftopoisomerase II. J Biol Chem 1999;274:17226–17235.
231. (a) Rosenberg B, van Camp L, Krigas T. Inhibition of cell division in Escherichia coli by electrolysis products from a platinum electrode. Nature 1965;205:698–699. (b) Rosenberg B, van Camp L, Trosko JE,Mansour VH. Platinum compounds: A new class of potent antitumour agents. Nature 1969;222:385–386.
232. (a) Prestayko AW, Crooke ST, Carter SK. Cisplatin, current status and new developments. New York: Academic; 1980. (b) Rosenberg B. Clinical aspects of platinum anticancer drugs. Metal Ions Biol Syst1980;11:127–196. (c) Rosenberg B. Fundamental studies with cisplatin. Cancer 1985;55:2303–2316. (d)Mascaux C, Paesmans M, Berghmans T, Branle F, Lafitte JJ, Lemaıˆtre F, Meert AP, Vermylen P, Sculier JP.
A systematic review of the role of etoposide and cisplatin in the chemotherapy of small cell lung cancer withmethodology assessment and meta-analysis. Lung Cancer 2000;30:23–36. (e) Sugiyama T, Yakushiji M,Noda K, Ikeda M, Kudoh R, Yajima A, Tomoda Y, Terashima Y, Takeuchi S, Hiura M, Saji F, Takahashi T,Umesaki N, Sato S, Hatae M, Ohashi Y. Phase II study of irinotecan and cisplatin as first-line chemotherapyin advanced or recurrent cervical cancer. Oncology 2000;58:31–37. (f) Smith IE, Talbot DC. Cisplatin andits analogues in the treatment of advanced breast cancer: A review. Brit J Cancer 1992;65:787–793. (g)McClay EF, Howell SB. A review: Intraperitoneal cisplatin in the management of patients with ovariancancer. Gynecol Oncol 1990;36:1–6.
233. (a) Lippard SJ. Metals in medicine. In: Bertini I, Gray HB, Lippard SJ, Valentine JS, editors. Bioinorganic chemistry. Mill Valley, CA: University Science Books; 1994. (b) Berners-Price SJ, Sadler PJ. Coordinationchemistry of metallodrugs: Insights into biological speciation from NMR spectroscopy. Coord Chem Rev1996;151:1–40. (c) Bloemink MJ, Reedijk J. Cisplatin and derived anticancer drugs: Mechanism andcurrent status of DNA binding. Met Ions Biol Syst 1996;32:641–685. (d) Wong E, Giandomenico CM.
Current status of platinum-based antitumor drugs. Chem Rev 1999;99:2451–2466. (e) Lippert B. Impact ofcisplatin on the recent development of Pt coordination chemistry: A case study. Coord Chem Rev1999;182:263–295. (f) Perez JM, Fuertes MA, Alonso C, Navarro-Ranninger C. Current status of thedevelopment of trans-platinum antitumor drugs. Crit Rev Oncol Hematol 2000;35:109–120. (g) Judson I,Kelland LR. New developments and approaches in the platinum arena. Drugs 2000;59:29–36. (h) Guo Z,Sadler PJ. Medicinal inorganic chemistry. Adv Inorg Chem 2000;49:183–306. (i) Cohen SM, Lippard SJ.
Cisplatin: From DNA damage to cancer chemotherapy. Prog Nucleic Acid Res 2001;67:93–130. (j) Sun H.
Metallodrugs. In: Grant DM, Harris RK, editors. Encyclopedia of nuclear magnetic resonance: A Sup-plement. Chichester, England: Wiley; 2002.
234. (a) Kauffman GB. Classics in coordination chemistry, Part 1. New York, NY: Dover; 1968. (b) Kauffman GB. Alfred Werner–Founder of Coordination Chemistry. Berlin: Springer-Verlag; 1966.
235. (a) Spingler B, Whittington DA, Lippard SJ. 2.4 A ˚ crystal structure of an oxaliplatin 1,2-d(GpG) intrastrand cross-link in a DNA dodecamer duplex. Inorg Chem 2001;40:5596–5602. (b) Coste F, Malinge JM, SerreL, Shepard W, Roth M, Leng M, Zelwer C. Crystal structure of a double-stranded DNA containing acisplatin interstrand cross-link at 1.63 A ˚ resolution: Hydration at the platinated site. Nucleic Acids Res 1999;27:1837–1846. (b) Takahara PM, Rosenzweig AC, Frederick CA, Lippard SJ. Crystal structure ofdouble-stranded DNA containing the major adduct of the anticancer drug cisplatin. Nature 1995;377:649–652. (c) Coll M, Sherman SE, Gibson D, Lippard SJ, Wang AH. Molecular structure of the complex formedbetween the anticancer drug cisplatin and d(pGpG):C222(1) crystal form. J Biomol Struct Dynam1990;8:315–330.
236. (a) Gelasco A, Lippard SJ. NMR solution structure of a DNA dodecamer duplex containing a cis- diammineplatinum(II) d(GpG) intrastrand cross-link, the major adduct of the anticancer drug cisplatin.
Biochemistry 1998;37:9230–9239. (b) Parkinson JA, Chen Y, del Socorro Murdoch P, Guo Z, Berners-Price SJ, Brown T, Sadler PJ. Sequence-dependent bending of DNA induced by cisplatin: NMR structuresof an A.T-rich 14-mer duplex. Chemistry (Germany) 2000;6:3636–3644. (c) Yang D, van Boom SS,Reedijk J, van Boom JH, Wang AH. Structure and isomerization of an intrastrand cisplatin-cross-linkedoctamer DNA duplex by NMR analysis. Biochemistry 1995;34:12912–12920.
237. Farrell N. Current status of structure–activity relationships of platinum anticancer drugs: Activation of the trans geometry. Met Ions Biol Syst 1996;32:603–639.
238. (a) Zou Y, van Houten B, Farrell N. Ligand effects on platinum binding to DNA. A comparison of DNA binding properties for cis- and trans-[PtCl2(amine)2] (amine ¼ NH3, pyridine). Biochemistry 1993;32:9632–9638. (b) Colombier C, Lippert B, Leng M. Interstrand cross-linking reaction in triplexes containinga monofunctional transplatin-adduct. Nucleic Acid Res 1996;24:4519–4524. (c) Muller J, Drumm M,Boudvillain M, Leng M, Sletten E, Lippert B. Parallel-stranded DNA with Hoogsteen base pairingstabilized by a trans-[Pt(NH3)(2)]2þ cross-link: Characterization and conversion into a homodimer and atriplex. J Biol Inorg Chem 2000;5:603–611.
239. (a) Farrell N, Kelland LR, Roberts JD, van Beusichem M. Activation of the trans geometry in platinum antitumor complexes: A survey of the cytotoxicity of trans complexes containing planar ligands in murineL1210 and human tumor panels and studies on their mechanism of action. Cancer Res 1992;52:5065–5072.
(b) van Beusichem M, Farrell N. Activation of the trans geometry in platinum antitumor complexes.
Synthesis, characterization, and biological activity of complexes with the planar ligands pyridine, N-methylimidazole, thiazole, and quinoline. Crystal and molecular structure of trans-dichlorobis(thiazole)-platinum(II). Inorg Chem 1992;31:634–639.
240. (a) Coluccia M, Nassi A, Loseto F, Boccarelli A, Mariggio MA, Gordano D, Intini F, Caputo P, Natile G. A trans-platinum complex showing higher antitumor activity than the cis congeners. J MedChem 1993;36:510–512. (b) Cini R, Caputo PA, Intini FP, Natile G. Mechanistic and stereo-chemical investigation of imino ethers formed by alcoholysis of coordinated nitriles: X-ray crystalstructures of cis- and trans-bis(1-imino-1-methoxyethane)dichloroplatinum(II). Inorg Chem 1995;34:1130–1137.
241. Murugkar A, Unnikrishnan B, Padhye S, Bhonde R, Teat S, Triantafillou E, Sinn E. Hormone anchored metal complexes. 1. Synthesis, structure, spectroscopy, and in vitro antitumor activity of testosteroneacetate thiosemicarbazone and its metal complexes. Metal-Based Drugs 1999;6:177–182.
242. Ivanov AI, Christofoulou J, Parkinson JA, Barnham KJ, Tucker A, Woodrow J, Sadler PJ. Cisplatin binding sites on human albumin. J Biol Chem 1998;273:14721–14730.
243. Cox MC, Barnham KJ, Frenkiel TA, Hoeschele JD, Mason AB, He Q-Y, Woodrow RC, Sadler PJ.
Identification of platination sites on human serum transferrin using C-13 and N-15 NMR spectroscopy.
J Biol Inorg Chem 1999;4:621–631.
244. Jamieson ER, Lippard SJ. Structure, recognition, and process of cisplatin–DNA adducts. Chem Rev 245. Weiss RB, Christian MC. New cisplatin analogues in development. A review. Drugs 1993;46:360–377.
246. Lebwohl D, Canetta R. Clinical development of platinum complexes in cancer therapy: An historical perspective and an update. Eur J Cancer 1998;34:1522–1534.
247. (a) Farrell N. Nonclassical platinum antitumor agents: Perspectives for design and development of new drugs complementary to cisplatin. Cancer Invest 1993;11:578–589. (b) Farrell N, Qu Y, Hacker MP.
Cytotoxicity and antitumor activity of bis(platinum) complexes. A novel class of platinum complexes active in cell lines resistant to both cisplatin and 1,2-diaminocyclohexane complexes. J Med Chem1990;33:2179–2184. (c) Farrell N, Qu Y, Feng L, Van Houten B. Comparison of chemical reactivity,cytotoxicity, interstrand cross-linking, and DNA sequence specificity of bis(platinum) complexescontaining monodentate or bidentate coordination spheres with their monomeric analogues. Biochemistry1990;29:9522–9531. (d) Jansen BAJ, van der Zwan J, den Dulk H, Brouwer J, Reedijk J. Dinuclearalkyldiamine platinum antitumor compounds: A structure –activity relationship study. J Med Chem 2001;44:245–249.
248. Recent studies: (a) Qu Y, Rauter H, Fontes APS, Bandarage R, Kelland L, Farrell N. Synthesis, characterization, and cytotoxicity of trifunctional dinuclear platinum complexes: Comparison of effects ofgeometry and polyfunctionality on biological activity. J Med Chem 2000;43:3189–3192. (b) Hofr C,Farrell N, Brabec V. Thermodynamic properties of duplex DNA containing a site-specific d(GpP)intrastrand crosslink formed by an antitumor dinuclear platinum complex. Nucleic Acids Res2001;29:2034–2040. (c) Cox JW, Berners-Price SJ, Davies MS, Qu Y, Farrell N. Kinetic analysis of thestepwise formation of a long-range DNA interstrand cross-link by a dinuclear platinum antitumor complex:Evidence for aquated intermediates and formation of both kinetically and thermodynamically controlledconformers. J Am Chem Soc 2001;123:1316–1326.
249. (a) Brabec V, Kasparkova J, Vrana O, Novakova O, Cox JW, Qu Y, Farrell N. DNA modifications by a novel bifunctional trinuclear platinum Phase I anticancer agent. Biochemistry 1999;38:6781–6790. (b) ColellaG, Pennati M, Bearzatto A, Leone R, Colangelo D, Manzotti C, Daidone MG, Zaffaroni N. Activity of atrinuclear platinum complex in human ovarian cancer cell lines sensitive and resistant to cisplatin:Cytotoxicity and induction and gene-specific repair of DNA lesions. Brit J Cancer 2001;84:1387–1390. (c)Servidei T, Ferlini C, Riccardi A, Meco D, Scambia G, Segni G, Manzotti C, Riccardi R. The noveltrinuclear platinum complex BBR3464 induces a cellular response different from cisplatin. Eur J Cancer2001;37:930–938.
250. (a) Davies MS, Thomas DS, Hegmans A, Berners-Price SJ, Farrell N. Kinetic and equilibria studies of the aquation of the trinuclear platinum phase II anticancer agent [{trans-PtCl(NH3)2}2{m-trans-Pt(NH3)2(NH2(CH2)6NH2)2}]4þ (BBR3464). Inorg Chem 2002;41:1101–1109. (b) McGregor TD,Hegmans A, Kasparkova J, Neplechova K, Novakova O, Penazova H, Vrana O, Brabec V, Farrell N. Acomparison of DNA binding profiles of dinuclear platinum compounds with polyamine linkers and thetrinuclear platinum phase II clinical agent BBR3464. J Biol Inorg Chem 2002;7:397–404.
251. (a) Weiss RB, Christian MC. New cisplatin analogues in development. A review. Drugs 1993; 46:360–377. (b) Christian MC. The current status of new platinum analogs. Seminars Oncol 1992;19:720–733. (c) Gordon M, Hollander S. Review of platinum anticancer compounds. J Med 1993;24:209–265.
252. (a) Braddock PD, Connors TA, Jones M, Khokhar AR, Melzack DH, Tobe ML. Structure and activity relationships of platinum complexes with anti-tumour activity. Chem-Biol Interact 1975;11:145–161. (b)Wong WS, Tindall VR, Wagstaff J, Bramwell V, Crowther D. Primary carcinoma of the fallopian tube:Favorable response to new chemotherapeutic agent, CHIP. J Roy Soc Med 1985;78:203–206. (c) BramwellVHC, Crowther D, O'Malley S, Swindell R, Johnson R, Cooper EH, Thatcher N, Howell A. Activity of JM9in advanced ovarian cancer: A phase I–II trial. Cancer Treat Rep 1985;69:409–416.
253. (a) Northcott SE, Marr JG, Secreast SL, Han F, Dezwaan J. Solution chemistry and analytical characterization of ormaplatin. J Pharm Biomed Anal 1991;9:1009–1018. (b) Tutsch KD, ArzoomanianRZ, Alberti D, Tombes MB, Feierabend C, Robins HI, Spriggs DR, Wilding G. Phase I clinical andpharmacokinetic study of an one-hour infusion of ormaplatin (NSC 363812). Invest New Drugs1999;17:63–72. (c) Luo FR, Wyrick SD, Chaney SG. Comparative neurotoxicity of oxaliplatin,ormaplatin, and their biotransformation products utilizing a rat dorsal root ganglia in vitro explant culturemodel. Cancer Chemother Pharmacol 1999;44:29–38.
254. (a) McKeage MJ, Raynaud F, Ward J, Berry C, O'Dell D, Kelland LR, Murrer B, Santaba´rabara P, Harrap KR, Judson IR. Phase I and pharmacokinetic study of an oral platinum complex given daily for 5 days inpatients with cancer. J Clin Oncol 1997;15:2691–2700. (b) Kelland LR. An update on satraplatin: The firstorally available platinum anticancer drug. Expert Opin Invest Drugs 2000;9:1373–1382.
255. Hall MD, Hambley TW. Platinum(IV) antitumour compounds: Their bioinorganic chemistry. Coord Chem Rev 2002; published on line February 14, 2002.
256. (a) Farrall N. Tranistion metal complexes as drugs and chemotherapeutic agents. Dordrecht, Netherlands: Kluwer; 1989. (b) Clarke MJ, Zhu F, Frasca DR. Non-platinum chemotherapeutic metallopharmaceuticals.
Chem Rev 1999;99:2511–2533.
257. (a) Ko¨pf-Maier P, Ko¨pf H. Antitumor metallocenes: New developments and toxicologic features.
Anticancer Res 1986;6:227–233. (b) Ko¨pf-Maier P, Ko¨pf H. Non-platinum group metal antitumor agents.
History, current status, and perspectives. Chem Rev 1987;87:1137–1152. (c) Ko¨pf-Maier P, Ko¨pf H.
Transition and main-group metal cyclopentadienyl complexes: Preclinical studies on a series of antitumoragents of different structure types. Struct Bond 1988;70:103–185.
258. (a) Christodoulou CV, Ferry DR, Fyfe DW, Young A, Doran J, Sheehan TM, Eliopoulos A, Hale K, Baumgart J, Sass G, Kerr DJ. Phase I trial of weekly scheduling and pharmacokinetics of titanocenedichloride in patients with advanced cancer. J Clin Oncol 1998;16:2761–2769. (b) Lu¨mmen G, Sperling H,Luboldt H, Otto T, Ru¨bben H. Phase II trial of titanocene dichloride in advanced renal-cell carcinoma.
Cancer Chemother Pharmacol 1998;42:415–417.
259. Kuo LY, Liu AH, Marks TJ. Metallocene interactions with DNA and DNA-processing enzymes. Met Ion Biol Syst 1996;33:53–85.
260. Harding MM, Mokdsi G, Mackay JP, Prodigalidad M, Lucas SW. Interactions of the antitumor agent molybdocene dichloride with oligonucleotides. Inorg Chem 1998;37:2432–2437.
261. Harding MM, Harden GJ, Field LD. A 31P NMR study of the interaction of the antitumor active metallocene Cp2MoCl2 with calf thymus DNA. FEBS Lett 1993;322:291–294.
262. Christodoulou CV, Eliopoulos AG, Young LS, Hodgkins L, Ferry DR, Kerr DJ. Anti-proliferative activity and mechanism of action of titanocene dichloride. Br J Cancer 1998;77:2088–2097.
263. Toney JH, Murthy MS, Marks TJ. Biodistribution and pharmacokinetics of vanadium following intraperitoneal administration of vanadocene dichloride to mice. Chem-Biol Inter 1985;56:45–54. (b)Murray JH, Harding MM. Organometallic anticancer agents: The effect of the central metal and halideligands on the interaction of metallocene dihalides Cp2MX2 with nucleic acid constituents. J Med Chem1994;37:1936–1941.
264. Moustatih A, Fiallo MML, Garnier-Suillerot A. Biofunctional antitumor compounds: Interaction of adriamycin with metallocene dichlorides. J Med Chem 1989;32:336–343.
265. (a) Kaji A, Ryoji M. Tetracycline. In: Hahn FE, editor. Antibiotics. Berlin: Springer; 1979. (b) Chopra I, Hawkey PM, Hinton M. Tetracycline, molecular and clinical aspects. J Antimicrob Chemother 1992;29:245–277.
266. Smythies JR, Benington F, Morin RD. On the molecular mechanism of action of the tetracyclines.
267. (a) Testa RT, Petersen PJ, Jacobus NV, Sum PE, Lee VJ, Tally FP. In vitro and in vivo antibacterial activities of the glycylcyclines, a new class of semisynthetic tetracyclines. Antimicrob Agen Chemother1993;37:2270–2277. (b) Sum PE, Lee VJ, Testa RT, Hlavka JJ, Ellestad GA, Bloom JD, Gluzman Y,Tally FP. Glycylcyclines. 1. A new generation of potent antibacterial agents through modification of9-aminotetracyclines. J Med Chem 1994;37:184–188. (c) Someya Y, Yamaguchi A, Sawai T. A novelglycylcycline, 9-(N,N-dimethylglycylamido)-6-demethyl-6-deoxytetracycline, is neither transportednor recognized by the transposon Tn10-encoded metal-tetracycline/H þ antiporter. Antimicrob AgenChemother 1995;39:247–249.
268. Chopra I. Glycylcyclines: Third-generation tetracycline antibiotics. Curr Opin Pharmacol 2001;1:464–469.
269. Johnson AP. GAR-936. Curr Opin Anti-Infect Invest Drugs 2000;2:164–170.
270. Mikulski CM, Fleming J, Fleming D. Chelates of tetracycline with first row transition metal perchlorates.
Inorg Chim Acta 1988;144:9–16.
271. (a) Ohyama T, Cowan JA. Calorimetric studies of metal binding to tetracycline. Role of solvent structure in defining the selectivity of metal ion–drug interactions. Inorg Chem 1995;34:3083–3086. (b) Nova´k-Pe´kliM, Mesbah ME-H, Petho¨ G. Equilibrium studies on tetracycline–metal ion system. J Pharm Biomed Anal1996;14:1025–1029. (c) Lambs L, Berthon G. Metal–ion tetracycline interactions in biological fluid. Part7. Quantitative investigation of methacycline complexes with Ca(II), Mg(II), Cu(II), and Zn(II) ions andassessment of their biological significance. Inorg Chim Acta 1988;151:33–43. (d) Wessels JM, Ford WE,Szymczak W, Schneider S. The complexation of tetracycline and anhydrotetracycline with Mg2þ andCa2þ: A spectroscopic study. J Phys Chem B 1998;102:9323–9331.
272. de Paula FCS, Carvalho S, Duarte HA, Paniago EB, Mangrich AS, Pereira-Maia EC. A physicochemical study of the tetracycline coordination to oxovanadium(IV). J Inorg Biochem 1999;76:221–230.
273. Jezowska-Bojczuk M, Lambs L, Kozlowski H, Berthon G. Metal ion–tetracycline interactions in biological fluids. 10. Structural investigations on copper(II) complexes of tetracycline, oxytetracycline,chlortetracycline, 4-(dedimethylamino)tetracycline, and 6-desoxy-6-demethyltetracycline and discussionof their binding modes. Inorg Chem 1993;32:428–437.
274. (a) Machado FC, Demicheli C, Garnier-Suillerot A, Beraldo H. Metal complexes of anhydrotetracycline. 2.
Absorption and circular dichroism study of Mg(II), Al(III), and Fe(III) complexes. Possible influence ofthe Mg(II) complex on the toxic side effects of tetracycline. J Inorg Biochem 1995;60:163–173. (b) DosSantos HF, Xavier ES, Zerner MC, De Almeida WB. Spectroscopic investigation of the Al(III)-anhydrotetracycline complexation process. J Mol Struc (Theochem) 2000;527:193–202. (c) Berthon G,Brion M, Lambs L. Metal ion-tetracycline interactions in biological fluids. 2. Potentiometric study of magnesium complexes with tetracycline, oxytetracycline, doxycycline, and minocycline, and discussion oftheir possible influence on the bioavailability of these antibiotics in blood plasma. J Inorg Biochem1983;19:1–18. (c) Brion M, Berthon G, Fourtillan JB. Metal ion–tetracyclines interactions in biologicalfluids. Potentiometric study of calcium complexes with tetracycline, oxytetracycline, doxycycline, andminocycline and simulation of their distributions under physiological conditions. Inorg Chim Acta1981;55:47–56.
275. (a) Jogun KH, Stezowski JJ. Chemical-structural properties of tetracycline derivatives. 2. Coordination and conformational aspects of oxytetracycline metal ion complexation. J Am Chem Soc 1976;98:6018–6026.
(b) Celotti M, Fazakerley GV. Conformation of various tetracycline species determined with aid of anuclear magnetic resonance relaxation probe. J Chem Soc Perkin II 1977;1319–1322. (c) Schnarr M,Matthies M, Lohmann W. The influence of different solvent on the interaction between metal ions andtetracycline. Z Naturforsch 1979;34:1156–1161.
276. White JP, Cantor CR. Role of magnesium in the binding of tetracycline to Escherichia coli ribosomes. J Mol 277. Berens C. Tetracyclines and RNA. In: Schroeder R, Wallis MG, editors. RNA-binding antibiotics. Landes: Bioscience; 2001.
278. Day LE. Tetracycline inhibition of cell-free protein synthesis II. Effect of the binding of tetracycline to the components of the system. J Bacteriol 1966:92:197–203.
279. (a) Fey G, Reiss M, Kersten H. Interaction of tetracyclines with ribosomal subunits from Escherichia coli.
A fluorometric investigation. Biochemistry 1973;12:1160–1164. (b) Bergeron J, Ammirati M, Danley D,James L, Norcia M, Retsema J, Strick CA, Su WG, Sutcliffe J, Wondrack L. Glycylcyclines bind to thehigh-affinity tetracycline ribosomal binding site and evade TetM and TetO-mediated ribosomal protection.
Antimicrob Agen Chemother 1996;40:2226–2228.
280. (a) Moazed D, Noller HF. Interaction of antibiotics with functional sites in 16S ribosomal RNA. Nature 1987;327:389–394. (b) Oehler R, Polacek N, Steiner G, Barta A. Interaction of tetracycline with RNA:Photo-incorporation into ribosomal RNA of Escherichia coli. Nucleic Acids Res 1997;25:1219–1224. (c)Ross JI, Eady EA, Cove JH, Cunliffe WJ. 16S rRNA mutation associated with tetracycline resistance in aGram-positive bacterium. Antimicrob Agen Chemother 1998;42:1702–1705.
281. (a) Brodersen DE, Clemons WM, Jr., Carter AP, Morgan-Warren RJ, Wimberly BT, Ramakrishnan V. The structural basis for the action of the antibiotics tetracycline, pactamycin, and hygromycin B on the 30Sribosomal subunit. Cell 2000;103:1143–1154. (b) Pioletti M, Schlunzen F, Harms J, Zarivach R,Gluhmann M, Avila H, Bashan A, Bartels H, Auerbach T, Jacobi C, Hartsch T, Yonath A, Franceschi F.
Crystal structures of complexes of the small ribosomal subunit with tetracycline, edeine and IF3. EMBO J2001;20:1829–1839.
282. (a) Liu Y, Tidwell RR, Leibowitz MJ. Inhibition of in vitro splicing of a group I intron of Pneumocystis carinii. J Eur Microbiol 1994;41:31–38. (b) Murray JB, Arnold JR. Antibiotic interactions with thehammerhead ribozyme: Tetracyclines as a new class of hammerhead inhibitor. Biochem J 1996;317:855–860.
283. (a) Speer BS, Shoemaker NB, Salyers AA. Bacterial resistance to tetracycline: Mechanism, transfer, and clinical significance. Clin Microbiol Rev 1992;5:387–399. (b) Schnappinger D, Hillen W. Tetracyclines:Antibiotic action, uptake, and resistance mechanisms. Arch Microbiol 1996;165:359–369. (c) Neu HC.
The crisis on antibiotic resistance. Science 1992;29:245–277.
284. (a) Yamaguchi A, Udagawa T, Sawai T. Transport of divalent cations with tetracycline as mediated by the transposon Tn10-encoded tetracycline resistance protein. J Biol Chem 1990;265:4089–4813. (b) HellenW, Berens C. Mechanisms under-lying expression of Tn10 encoded tetracycline resistance. Ann RevMicrobiol 1994;48:345–369.
285. (a) Takahashi M, Altschmied L, Hillen W. Kinetic and equilibrium characterization of the Tet repressor- tetracycline complex by fluorescence measurements. Evidence for divalent metal ion requirement andenergy transfer. J Mol Biol 1986;187:341–348. (b) Orth P, Saenger W, Hinrichs W. Tetracycline-chelatedMg2þ ion initiates helix unwinding in Tet repressor induction. Biochemistry 1999;38:191–198.
286. Orth P, Cordes F, Schnappinger D, Hillen W, Saenger W, Hinrichs W. Conformational changes of the Tet repressor induced by tetracycline trapping. J Mol Biol 1998;279:439–447.
287. (a) Hinrichs W, Kisker C, Du¨vel M, Mu¨ller A, Tovar K, Hillen W, Saenger W. Structure of the Tet repressor- tetracycline complex and regulation of antibiotic resistance. Science 1994;264:418–420. (b) Kisker C,Hinrichs W, Tovar K, Hillen W, Saenger W. The complex formed between Tet repressor and tetracycline-Mg2þ reveals mechanism of antibiotic resistance. J Mol Biol 1995;247:260–280. (c) Orth P, SchnappingerD, Sum PE, Ellestad GA, Hillen W, Saenger W, Hinrichs W. Crystal structure of the tet repressor in complexwith a novel tetracycline, 9-(N,N-dimethylglycylamido)-6-demethyl-6-deoxy-tetracycline. J Mol Biol1999;285:455–461.
288. (a) Ettner N, Hillen W, Ellestad GA. Enhanced site-specific cleavage of the tetracycline repressor by tetracycline complexed with iron. J Am Chem Soc 1993;115:2546–2548. (b) Ettner N, Metzger JW,Lederer T, Hulmes JD, Kisker C, Hinrichs W, Ellestad GA, Hillen W. Proximity mapping of the Tetrepressor-tetracycline-Fe2þ complex by hydrogen peroxide mediated protein cleavage. Biochemistry1995;34:22–31.
289. Brewer GA, Florey K, editors. Analytical profiles of drug substances (No. 9 in series) New York: Academic; 290. Hanson DJ. Human health effects of animal feed drugs unclear. Chem Eng News 1985;63:7–11.
291. Arky R. Physicians' desk reference for nonprescription drugs, 18th edn. Montvale, NJ: Medical Economics Company; 1997.
292. Ming L-J, Epperson JD. Metal binding and structure–activity relationship of the metalloantibiotic peptide bacitracin. J Inorg Biochem 2002;91:46–58.
293. (a) Morris M. Primary structural confirmation of components of the bacitracin complex. Biol Mass Spect 1994;23:61–70. (b) Siegel MM, Huang J, Lin B, Tsao R. Structures of bacitracin A and isolated congeners:Sequencing of cyclic peptides with blocked linear side chains by electrospray ionization massspectrometry. Biol Mass Spect 1994;23:196–204.
294. (a) Pfaender P, Specht D, Heinrich G, Schwarz E, Kuhn E, Simlot MM. Enzymes of Bacillus licheniformis in the biosynthesis of bacitracin A. FEBS Lett 1973;32:100–104. (b) Ishihara H, Shimura K. Biosynthesisof bacitracins. Biochim Biophys Acta 1974;338:588–600. (c) Frøyshov Ø, Laland SG. On the biosynthesisof bacitracin by a soluble enzyme complex from B. licheniformis. Eur J Biochem 1974;42:235–242. (d)Rieder H, Heinrich G, Breuker E, Simlot MM, Pfaender P. Bacitracin synthetase. Meth Enzymol1975;43:548–559. (e) Frøyshov Ø. The production of bacitracin synthetase by Bacillus licheniformisATCC 10716. FEBS Lett 1977;81:315–318. (f) Ogawa I, Ishihara H, Shimura K. Component I protein ofbacitracin synthetase: A multifunctional protein. FEBS Lett 1981;124:197–201. (g) Podlesek Z, GrabnarM. Genetic mapping of the bacitracin synthetase gene(s) in Bacillus licheniformis. J Gen Microbiol1987;133:3093–3097. (h) Ishihara H, Hara N, Iwabuchit T. Molecular-cloning and expression inEscherichia coli of the Bacillus licheniformis bacitracin synthesase-2 gene. J Bacteriol 1989;171:1705–1711. (i) Konz D, Klens A, Scho¨rgendorfer K, Marahiel MA. The bacitracin biosynthesis operon ofBacillus licheniformis ATCC 10716: Molecular characterization of three multi-modular peptidesynthetases. Chem Biol 1997;4:927–937.
295. Frøyshov Ø. Enzyme-bound intermediates in the biosynthesis of bacitracin. Eur J Biochem 1975;59:201– 296. (a) Ishihara H, Shimura K. Thiazoline ring formation in bacitracin biosynthesis. FEBS Lett 1979;99:109– 112. (b) Ishihara H, Shimura K. Further evidence for the presence of a thiazoline ring in theisoleucylcysteine dipeptide intermediate in bacitracin biosynthesis. FEBS Lett 1988;226:319–323.
297. (a) Guilvout I, Mercereau-Puijalon O, Bonnefoy S, Pugsley AP, Carniel E. High-molecular-weight protein 2 of Yersinia enterocolitica is homologous to AngR of Vibrio anguillarum and belongs to a family ofproteins involved in nonribosomal peptide synthesis. J Bacteriol 1993;175:5488–5504. (b) Tolmasky ME,Actis LA, Crosa JH. A single amino acid change in AngR, a protein encoded by pJM1-like virulenceplasmids, results in hyperproduction of anguibactin. Infect Immun 1993;61:3228–3233.
298. Frøyshov Ø, Mathiesen A, Haavik HI. Regulation of bacitracin synthetase by divalent metal ions in Bacillus licheniformis. J Gen Microbiol 1980;117:163–167.
299. Eppelmann K, Doekel S, Marahiel MA. Engineered biosynthesis of the peptide antibiotic bacitracin in the surrogate host Bacillus subtilis. J Biol Chem 2001;276:34824–34831.
300. Storm DR, Strominger JL. Binding of bacitracin to cells and protoplasts of Micrococcus lysodeikticus.
J Biol Chem 1974;249:1823–1827.
301. (a) Scogin DA, Mosberg HI, Storm DR, Gennis RB. Binding of nickel and zinc ions to bacitracin A.
Biochemistry 1980;19:3342–3348. (b) Seebauer EG, Duliba EP, Scogin DA, Gennis RB, Belford RL. EPRevidence on the structure of the copper(II)-bacitracin A complex. J Am Chem Soc 1983;105:4926–4929.
(c) Mosberg HI, Scogin DA, Storm DR, Gennis RB. Proton nuclear magnetic resonance studies onbacitracin A and its interaction with zinc ion. Biochemistry 1980;19:3353–3357.
302. Storm DR, Strominger JL. Complex formation between bacitracin peptides and isoprenyl pyrophosphates.
J Biol Chem 1973;248:3940–3945. (b) Stone KJ, Strominger JL. Mechanism of action of bacitracin:Complexation with metal ion and C55-isoprenyl pyrophosphate. Proc Nat Acad Sci USA 1971;68:3223–3227.
303. (a) Higashi Y, Siewert G, Strominger JL. Biosynthesis of the peptidoglycan of bacterial cell walls. XIX.
Isoprenoid alcohol phosphokinase. J Biol Chem 1970;245:3683–3690. (b) Bhagavan NV. MedicalBiochemistry, 4th edn. Chapter 16. San Diego: Harcourt; 2001.
304. Drabløs F, Nicholson DG, Rønning M. EXAFS study of zinc coordination in bacitracin A. Biochim Biophys Acta 1999;1431:433–442.
305. Epperson JD, Ming L-J. Proton NMR studies of Co(II) complexes of bacitracin analogous: Insight into structure–activity relationship. Biochemistry 2000;39:4037–4045.
306. Epperson JD. Paramagnetic cobalt(II) as a nuclear magnetic resonance probe for the study of metallo- macromolecules: From peptides and proteins to dendrimers. Ph.D. Dissertation, University of SouthFlorida, 1999.
307. Pons M, Feliz M, Molins MA, Giralt E. Conformational analysis of bacitracin A a naturally occurring lariat.
Biopolymers 1991;31:605–612. (b) Kobayashi N, Takenouchi T, Endo S, Munekata E. 1H NMR study ofthe conformation of bacitracin A in aqueous solution. FEBS Lett 1992;305:105–109.
308. (a) Pfeffer S, Hohne W, Branner S, Wilson K, Betzel C. X-ray structure of the antibiotic bacitracin A. FEBS Lett 1991;285:115–119. (b) Pfeffer-Hennig S, Dauter Z, Hennig M, Hohne W, Wilson K, Betzel C.
Three dimensional structure of the antibiotic bacitracin A complexed to two different subtilisin proteases:Novel mode of enzyme inhibition. Adv Exp Med Biol 1996;379:29–41.
309. Examples are found in (a) Gehm BD, Rosner MR. Regulation of insulin, epidermal growth factor, and transforming growth factor-alpha levels by growth factor-degrading enzymes. Endocrinology1991;128:1603–1610. (b) Mantle D, Lauffart B, Gibson A. Purification and characterization of leucylaminopeptidase and pyroglutamyl aminopeptidase from human skeletal muscle. Clin Chim Acta Int J ClinChem 1991;197:35–45. (c) Janas J, Sitkiewicz D, Warnawin K, Janas RM. Characterization of a novel,high molecular weight, acidic, endothelin-1 inactivating metalloendopeptidase from the rat kidney.
J Hypertension 1994;12:1155–1162.
310. (a) Essex DW, Li M, Miller A, Feinman RD. Protein disulfide isomerase and sulfhydryl-dependent pathways in platelet activation. Biochemistry 2001;40:6070–6075. (b) Ta¨ger M, Kro¨ning H, Thiel U,Ansorge S. Membrane-bound proteindisulfide isomerase (PDI) is involved in regulation of surfaceexpression of thiols and drug sensitivity of B-CLL cells. Exp Hematol 1997;25:601–607. (c) Clive DR,Greene JJ. Association of protein disulfide isomerase activity and the induction of contact inhibition.
Exp Cell Res 1994;214:139–144. (d) Mandel R, Ryser HJ, Ghani F, Wu M, Peak D. Inhibition of areductive function of the plasma membrane by bacitracin and antibodies against protein disulfide-isomerase. Proc Nat Acad Sci USA 1993;90:4112–4116. (d) Mizunaga T, Katakura Y, Miura T, MaruyamaY. Purification and characterization of yeast protein disulfide isomerase. J Biochem 1990;108:846–851.
311. Mou Y, Ni H, Wilkins JA. The selective inhibition of beta 1 and beta 7 integrin-mediated lymphocyte adhesion by bacitracin. J Immunol 1998;161:6323–6329.
312. Easwaran KRK. Interaction between valinomycin and metalions. Metal Ions Biol Syst 1985;19:109–137.
313. Steinrauf LK. Beauvericin and the other enniatins. Metal Ions Biol Syst 1985;19:139–171.
314. Nawata Y, Ando K, Iitaka Y. Nactins: Their complexes and biological properties. Metal Ions Biol Syst 315. Painter GR, Pressman BC. Cation complexes of the monovalent and polyvalent carboxylic ionophores: Lasalocid (X-537A), monensin, A23187 (calcimycin), and related antibiotics. Metal Ions Biol Syst1985;19:229–294.
316. Dobler M. Ionophores and their structures. New York: Wiley; 1981.
317. (a) Neilands JB, Valenta JR. Iron-containing antibiotics. Metal Ions Biol Syst 1985;19:313–333. (b) Neilands JB. Siderophores. Adv Inorg Biochem 1983;5:137–199. (c) Drechsel H, Jung G. Peptidesiderophores. J Pept Sci 1998;4:147–181.
318. (a) Song MK, Choi SH. Growth promoters and their effects on beef production—Review. Asian-Aust J Animal Sci 2001;14:123–135. (b) Popp JD, McCaughey WP, Cohen RDH, McAllister TA, Majak W.
Enhancing pasture productivity with alfalfa: A review. Can J Plant Sci 2000;80:513–519. (c) Duffield TF,Bagg RN. Use of ionophores in lactating dairy cattle: A review. Can Vet J 2000;41:388–394.
319. Leevy WM, Donato GM, Ferdani R, Goldman WE, Schlesinger PH, Gokel GW. Synthetic hydraphile channels of appropriate length kill Escherichia coli. J Am Chem Soc 2002;124:9022–9023.
320. (a) Tishchenko GN, Karimov Z. Structure of Rb complex of LDLLDL analog of enniatin B in crystals.
Crystallography (Russia) 1978;23:729–742. (b) Dobler M, Dunitz JD, Krajewski J. Structure of the Kþcomplex with enniatin B, a macrocyclic antibiotic with K þ transport properties. J Mol Biol 1969;42:603–606. (c) Tishchenko GN, Zhukhlistova NE. Crystal and molecular structure of the membrane-activeantibiotic enniatin C. Crystallogr Rep 2000;45:619–625.
321. (a) Dobler M. The crystal structure of nonactin. Helv Chim Acta 1972;55:1371–1384. (b) Dobler M, Phizackerley RP. The crystal structure of the NaNCS complex of nonactin. Helv Chim Acta 1974;57:664–674. (c) Iitaka Y, Sakamaki T, Nawata Y. Molecular structures of tetranactin and its alkali metal ioncomplexes. Chem Lett 1972;1225–1230. (d) Nawata Y, Sakamaki T, Iitaka Y. The crystal and molecularstructures of tetranactin. Acta Crystallogr B 1974;B30:1047–1053.
322. (a) Hamilton JA, Sabesan MN, Steinrauf LK. Crystal structure of valinomycin potassium picrate: Anion effects on valinomycin cation complexes. J Am Chem Soc 1981;103:5880–5885. (b) Steinrauf LK, Hamilton JA, Sabesan MN. Crystal structure of valinomycin-sodium picrate. Anion effects onvalinomycin-cation complexes. J Am Chem Soc 1982;104:4085–4091. (c) Smith GD, Duax WL, LangDA, de Titta GT, Edmonds JW, Rohrer DC, Weeks CW. Crystal and molecular structure of the triclinic andmonoclinic forms of valinomycin, C54H90N6O18. J Am Chem Soc 1975;97:7242–7247. (d) Devarajan S,Vijayan M, Easwaran KRK. Conformation of valinomycin in its barium perchlorate complex from X-raycrystallography and n.m.r. Int J Prot Pept Res 1984;23:324–333.
323. (a) Haynes DH, Pressman BC. X537A: A Ca2þ ionophore with a polarity-dependent and complexation- dependent fluorescence signal. J Membr Biol 1974;16:195–205. (b) Haynes DH, Pressman BC. Two-phasepartition studies of alkali cation complexation by ionophores. J Membr Biol 1974;18:1–21.
324. Eisenman G, Krasne S, Ciani S. Kinetic and equilibrium components of selective ionic permeability mediated by nactin-type and valinolycin-type carriers having systematically varied degrees of methylation.
Ann NY Acad Sci 1975;264:34–60.
325. Prestegard JH, Chan SI. Proton magnetic resonance studies of the cation-binding properties of nonactin. II.
Comparison of the sodium ion, potassium ion, and cesium ion complexes. J Am Chem Soc 1970;92:4440–4446.
326. (a) Ward DL, Wei KT, Hoogerheide JG, Popov AI. Crystal and molecular structure of sodium bromide complex of monensin, C 36H62O11 Naþ Br. Acta Cryst 1978;334:110 – 115. (b) Johnson SM, Herrin J, Liu SJ, Paul IC. Crystal and molecular structure of the barium salt of an antibiotic containing a high proportionof oxygen. J Am Chem Soc 1970;92:4428–4435.
327. Pankiewicz R, Schroeder G, Gierczyk B, Wojciechowski G, Brzezinski B, Bartl F, Zundel G. Li-7 NMR and FTIR studies of lithium, potassium, rubidium, and cesium complexes with ionophore lasalocid in solution.
Biopolymers 2001;62:173–182.
328. Schroeder G, Gierczyk B, Brzezinski B, Rozalski B, Bartl F, Zundel G, Sosnicki J, Grech E. Na-23 NMR and FT-IR studies of sodium complexes with the ionophore lasalocid in solution. J Mol Struc 2000;516:91–98.
329. (a) Mimouni M, Lyazghi R, Juillard J. Formation and structure of alkali metal, thallium, silver and alkaline- earth cation complexes with the ionophore lasalocid free acid form in methanol from NMR experiments.
New J Chem 1998;22:367–372. (b) Mimouni M, Hebrant M, Dauphin G, Juillard J. Monovalentcation salts of the bacterial ionophore monensin in methanol. Structural aspects from NMR experiments.
J Chem Res 1996;278–279. (c) Mimouni M, Malfreyt P, Lyazghi R, Palma M, Pascal Y, Dauphin G,Juillard J. Interactions between metal-cations and the ionophore lasalocid. 13. Structure of 1/1 and 2/1lasalocid anio-divalent cation complexes in methanol. J Chem Soc Perkin Trans II 1995;1939–1947.
330. (a) Chen ST, Springer CS. Interaction of antibiotic lasalocid A (X537A) with praseodymium(III) in methanol. Bioinorg Chem 1978;9:101–122. (b) Grandjeau J, Laszlo P. Synergistic transport ofpraseodymium(III) ion across lipid bilayers in the presence of two chemically distinct ionophores. J AmChem Soc 1984;106:1472–1476.
331. Tsukube H, Takeishi H, Yoshida Z. Recognition of metal complex guests: Supermolecular extraction of water-soluble lanthanide complexes by biological lasalocid ionophore. Inorg Chim Acta 1996;251:1–3.
332. Lindoy LF. Outer-sphere and inner-sphere complexation of cations by the natural ionophore lasalocid A.
Coord Chem Rev 1996;148:349–368.
333. Cross TA. Solid-state nuclear magnetic resonance characterization of gramicidin channel structure. Met 334. Koeppe RE II, Kimura M. Computer building of beta-helical polypeptide models. Biopolymers 1984; 335. (a) Myers VB, Haydon DA. Ion transfer across lipid membranes in the presence of gramicidin A. II. The ion selectivity. Biochim Biophys Acta 1972;274:313–322. (b) Bamberg E, Noda K, Gross E, La¨uger P. Single-channel parameters of gramicidin A, B, and C. Biochim Biophys Acta 1976;449:223–228. (c) Sawyer DB,Williams LP, Whaley WL, Koeppe RE II, Andersen OS. Gramicidins A, B, and C form structurallyequivalent ion channels. Biophys J 1990;58:1207–1212.
336. (a) Hladky SB, Haydon DA. Ion transfer across lipid membranes in the presence of gramicidin A. I. Studies of the unit conductance channel. Biochim Biophys Acta 1972;247:294–312. (b) Urry DW, Trapane TL,Walker JT, Prasad KU. On the relative lipid membrane permeability of Naþ and Ca2þ. A physical basis forthe messenger role of Ca2þ . J Biol Chem 1982;257:6659–6661.
337. (a) Wallace BA, Ravikumar K. The gramicidin pore: Crystal structure of a cesium complex.
Science 1988;241:182 – 187. (b) Wallace BA, Hendrickson WA, Ravikumar K. The use of single-wavelength anomalous scattering to solve the crystal structure of a gramicidin A/cesium chloridecomplex. Acta Crystallogr B 1990;46:440– 446. (c) Wallace BA, Janes RW. Co-crystals of gramicidinA and phospholipid. A system for studying the structure of a transmembrane channel. J Mol Biol 1991;217:625 – 627. (d) Doyle DA, Wallace BA. Crystal structure of the gramicidin/potassiumthiocyanate complex. J Mol Biol 1997;266:963 – 977. (e) Burkhart BM, Li N, Langs DA, PangbornWA, Duax WL. The conducting form of gramicidin A is a right-handed double-stranded doublehelix. Proc Nat Acad Sci USA 1998;95:12950– 12955. (f) Wallace BA. X-ray crystallographic struc-tures of gramicidin and their relation to the Streptomyces lividans potassium channel structure.
Novartis Found Symp 1999;225:23 – 32 and 33 – 37. (g) Doi M, Fujita S, Katsuya Y, Sasaki M,Taniguchi T, Hasegawa H. Antiparallel pleated beta-sheets observed in crystal structures of N,N-bis(trichloroacetyl) and N,N-bis(m-bromobenzoyl) gramicidin S. Arch Biochem Biophys 2001;395:85– 93.
338. (a) Townsley LE, Tucker WA, Sham S, Hinton JF. Structure of gramicidins A, B, and C incorporated into sodium dodecyl sulfate micelles. Biochemistry 2001;40:11674–11686. (b) Separovic F, Barker S,Delahunty M, Smith R. NMR structure of C-terminally tagged gramicidin channels. Biochim Biophys Acta1999;1416:48–56.
339. (a) Kim S, Quine JR, Cross TA. Complete cross-validation and R-factor calculation of a solid-state NMR derived structure. J Am Chem Soc 2001;123:7292–7298. (b) Fu R, Cotten M, Cross TA. Inter- and intra-molecular distance measurements by solid-state MAS NMR: Determination of gramicidin A channel dimerstructure in hydrated phospholipid bilayers. J Biomol NMR 2000;16:261–268. (c) Kovacs F, Quine J,Cross TA. Validation of the single-stranded channel conformation of gramicidin A by solid-state NMR.
Proc Nat Acad Sci USA 1999;96:7910–7915. (d) Quine JR, Brenneman MT, Cross TA. Protein structuralanalysis from solid-state NMR-derived orientational constraints. Biophys J 1997;72:2342–2348. (e)Ketchem RR, Hu W, Cross TA. High-resolution conformation of gramicidin A in a lipid bilayer by solid-state NMR. Science 1993;261:1457–1460.
340. (a) Koeppe RE II, Killian JA, Greathouse DV. Orientations of the tryptophan 9 and 11 side chains of the gramicidin channel based on deuterium nuclear magnetic resonance spectroscopy. Biophys J 1994;66:14–24. (b) Killian JA, Taylor MJ, Koeppe RE II. Orientation of the valine-1 side chain of the gramicidintransmembrane channel and implications for channel functioning. A 2H NMR study. Biochemistry1992;31:11283–11290. (c) Separovic F, Gehrmann J, Milne T, Cornell BA, Lin SY, Smith R. Sodium ionbinding in the gramicidin A channel. Solid-state NMR studies of the tryptophan residues. Biophys J1994;67:1495–1500. (d) Separovic F, Hayamizu K, Smith R, Cornell BA. C-13 chemical shift tensor ofL-tryptophan and its application to polypeptide structure determination. Chem Phys Lett 1991;181:157–162. (e) Smith R, Thomas DE, Separovic F, Atkins AR, Cornell BA. Determination of the structure of amembrane-incorporated ion channel. Solid-state nuclear magnetic resonance studies of gramicidin A.
Biophys J 1989;56:307–314. (c) Cornell BA, Separovic F, Baldassi AJ, Smith R. Conformation andorientation of gramicidin A in oriented phospholipid bilayers measured by solid state carbon-13 NMR.
Biophys J 1988;53:67–76.
341. Burkhart BM, Gassman RM, Langs DA, Pangborn WA, Duax WL, Pletnev V. Gramicidin D conformation, dynamics and membrane ion transport. Biopolymers 1999;51:129–144.
342. Sigel A, Sigel H. Iron transport and storage in microorganisms, plants, and animals. New York: Dekker; 343. New York, Joshi RR, Ganesh KN. Chemical cleavage of plasmid DNA by Cu(II), Ni(II), and Co(III) desferal complexes. Biochem Biophys Res Commun 1992;182:588–592. (b) Joshi RR, Ganesh KN. Spe-cific cleavage of DNA at CG sites by Co(III) and Ni(II) desferal complexes. FEBS Lett 1992;313:303–306.
344. Bagg A, Neilands JB. Molecular mechanism of regulation of siderophore-mediated iron assimilation.
Microbiol Rev 1989;51:509–518.
345. (a) Ecker DJ, Emery T. Iron uptake from ferrichrome A and iron citrate in Ustilago sphaerogena.
J Bacteriol 1983;155:616–622. (b) Wayne R, Frick K, Neilands JB. Siderophore protection against colicinsM, B, V, and Ia in Escherichia coli. J Bacteriol 1976;126:7–12.
346. Dhungana S, White PS, Crumbliss AL. Crystal structure of ferrioxamine B: A comparative analysis and implications for molecular recognition. J Biol Inorg Chem 2001;6:810–818.
347. Hossain MB, Jalal MAF, van der Helm D. The structure of ferrioxamine D1 ethanol water (1/2/1).
Acta Crystallogr C 1986;42:1305–1310.
348. (a) van der Helm D, Poling M. Crystal structure of ferrioxamine E. J Am Chem Soc 1976;98:82–86. (b) Hossain MB, Jalal MAF, van der Helm D, Shimizu K, Akiyama M. Crystal structure of retro-isomer of thesiderophore ferrioxamine E. J Chem Crystallogr 1998;28:53–56.
349. Hossain MB, van der Helm D, Poling M. The structure of deferriferrioxamine E (nocardamin), a cyclic trihydroxamate. Acta Crystallogr Sect B 1983;39:258–263.
350. (a) Norrestam R, Stensland B, Branden C-I. On the conformation of cyclic iron-containing hexapeptides: The crystal and molecular structure of ferrichrysin. J Mol Biol 1975;99:501–506. (b) van der Helm D,Baker JR, Eng-Wilmot DL, Hossain MB, Loghry RA. Crystal structure of ferrichrome and a comparison with the structure of ferrichrome A. J Am Chem Soc 1980;102:4224–4231. (c) Zalkin A, Forrester JD,Templeton DH. Ferrichrome A tetrahydrate. Determination of crystal and molecular structure. J Am ChemSoc 1966;88:1810–1814.
351. Knu¨sel F, Zimmermann W. Sideromycins. In: Corcoran JW, Hahn FE, editors. Antibiotics III. Mechanisms of action of antimicrobial and antitumor agents. Verlag, Heidelberg: Springer; 1975. pp 653–667.
352. Wayne R, Neilands JB. Evidence for common binding sites for ferrichrome compounds and bacteriophage phi 80 in the cell envelope of Escherichia coli. J Bacteriol 1975;121:497–503.
353. (a) Fecker L, Braun V. Cloning and expression of the fhu genes involved in iron(III)-hydroxamate uptake by Escherichia coli. J Bacteriol 1983;156:1301–1314. (b) Braun V, Hantke K, Stauder W. Identification of thesid outer membrane receptor protein in Salmonella typhimurium SL1027. Mol Gen Genet 1977;155:227–229.
354. Braun V, Gu¨nthner K, Hantke K, Zimmermann L. Intracellular activation of albomycin in Escherichia coli and Salmonella typhimurium. J Bacteriol 1983;156:308–315.
355. (a) Locher KP, Rees B, Koebnik R, Mitschler A, Moulinier L, Rosenbusch J, Moras D. Transmembrane signaling across the ligand-gated FhuA receptor: Crystal structure of free and ferrichrome-bound statesreveal allosteric changes. Cell 1998;95:45–56. (b) Ferguson AD, Hofmann E, Coulton JW, Diederichs K,Welte W. Siderophore-mediated iron transport: Crystal structure of FhuA with bound lipopolysaccharide.
Science 1998;282:2215–2220. (c) Ferguson AD, Braun V, Fiedler H-P, Coulton JW, Diederichs K, WelteW. Crystal structure of the antibiotic albomycin in complex with the outer membrane transporter FhuA.
Prot Sci 2000;9:956–963.
356. Clarke TE, Ku S-Y, Dougan DR, Vogel HJ, Tari LW. The structure of the ferric siderophore binding protein FhuD complexed with gallichrome. Nat Struct Biol 2000;7:287–291.
357. Ferguson AD, Ko¨dding J, Walker G, Bo¨s C, Coulton JW, Diederichs K, Braun V, Welte W. Active transport of an antibiotic rifamycin derivative by the outer-membrane protein FhuA. Structure 2001;9:707–716.
358. Buchanan SK, Smith BS, Venkatramani L, Xia D, Esser L, Palnitkar M, Chakraborty R, van der Helm D, Deisenhofer J. Crystal structure of the outer membrane active transporter FepA from Escherichia coli. NatStruct Biol 1999;6:56–63.
359. Ferguson AD, Chakraborty R, Smith BS, Esser L, van der Helm D, Deisenhofer J. Structural basis of gating by the outer membrane transporter FecA. Science 2002;295:1715–1719.
360. Buchanan SK. b-barrel proteins from bacterial outer membranes: Structure, function, and refolding.
Curr Opin Struct Biol 1999;9:455–461.
361. (a) Baslas RK, Zamani R, Nomani AA. Studies on the metal-complex of acetyl salicylic acid (aspirin).
Experientia 1979;35:455–456. (b) Gonzalez BE, Daeid NN, Nolan KB, Farkas E. Complex formationbetween transition metal ions and salicylglycine, a metabolite of aspirin. Polyhedron 1994;13:1495–1499.
(c) Nolan KB, Soudi AA. Synthesis and characterization of copper(II), zinc(II), and cobalt complexes ofsalicylglycine, a metabolite of aspirin. Inorg Chim Acta 1995;230:209–210. (d) Muller JG, Burrows CJ.
Metallodrug complexes that mediate DNA and lipid damage via sulfite autoxidation: Copper(II) famotidineand iron(III) bis(salicylglycine). Inorg Chim Acta 1998;275-276:314–319.
362. Abuhijleh AL. Mononuclear and binuclear copper(II) complexes of the anti-inflammatory drug ibuprofen: Synthesis, characterization, and catecholase-mimetic activity. J Inorg Biochem 1994;55:255–262.
363. Fini A, Feroci G, Fazio G. Interaction between indomethacin and heavy metal ions in aqueous solution.
Eur J Pharm Sci 2001;13:213–217.
364. Afanaseva IB, Ostrakhovitch EA, Mikhalchik EV, Ibragimova GA, Korkina LG. Enhancement of antioxidant and anti-inflammatory activities of bioflavonoid rutin by complexation with transition metals.
Biochem Pharmacol 2001;61:677–684.
365. (a) Kovala-Demertzi D, Mentzafos D, Terzis A. Metal complexes of the anti-inflammatory drug sodium [2-[(2,6-dichlorophenyl)amino]phenyl]acetate (diclofenac sodium). Molecular and crystal structure ofcadmium diclofenac. Polyhedron 1993;12:1361–1370. (b) Kovala-Demertzi D. Transition metalcomplexes of diclofenac with potentially interesting anti-inflammatory activity. J Inorg Biochem 2000;79:153–157.
366. Underhill AE, Bougourd SA, Flugge ML, Gale SE, Gomm PS. Metal complexes of anti-inflammatory drugs. Part VIII: Suprofen complex of copper(II). J Inorg Biochem 1993;52:139–144.
367. (a) Underhill AE, Bury A, Odling RJ, Fleet MB, Stevens A, Gomm PS. Metal complexes of antiinflammatory drugs. Part VII: Salsalate complex of copper(II). J Inorg Biochem 1989;37:1–5. (b)Bury A, Underhill AE, Fleet MB, Keymer PJ, Stevens A, Gomm PS. Metal complexes of anti-inflammatorydrugs. Part VI. 2-aminomethyl-4(1,1-dimethylethyl)-6-iodophenol (MK-447) complex of copper(II).
Inorg Chim Acta 1989;158:181–184. (c) Bury A, Underhill AE, Burke K, Fleming RPE, Davies JR, GommPS. Metal complexes of anti-inflammatory drugs. Part V. Meclofenamic acid complexes of manganese(II), iron(III), cobalt(II), nickel(II), copper(II), and zinc(II). Inorg Chim Acta 1988;152:171–175. (d) Bury A,Underhill AE, Kemp DR, O'Shea NJ, Smith JP, Gomm PS. Metal complexes of anti-inflammatory drugs.
Part IV. Tenoxicam complexes of manganese(II), iron(III), cobalt(II), nickel(II), and copper(II). InorgChim Acta 1987;138:85–89. (e) Underhill AE, Blundell RP, Gomm PS, Jacks ME. Metal complexes ofanti-inflammatory drugs. Part III. Nictindole complexes of copper(II) halides. Inorg Chim Acta1986;124:133–136. (f) Harrison DO, Thomas R, Underhill AE, Fletcher JK, Gomm PS, Hallway F.
Metal complexes of anti-inflammatory drugs–I. Complexes of iron(III), cobalt(II), nickel(II), copper(II),and zinc(II) with 4-hydroxy-3-(5-methyl-3-isoxazolocarbamyl)-2-methyl-2H-1,2-benzothiazine-1,1-dioxide (isoxicam). Polyhedron 1985;4:681–685.
368. Kirkova M, Atanassova M, Russanov E. Effects of cimetidine and its metal complexes on nitroblue tetrazolium and ferricytochrome c reduction by superoxide radicals. Gen Pharmacol 1999;33:271–276.
369. (a) Duda AM, Kowalik-Jankowska T, Kozlowski H, Kupka T. Histamine H-2 antagonists—Powerful ligands for copper(II)-reinterpretation of the famotidin-copper(II) system. J Chem Soc Dalton Trans 1995;2909–2913. (b) Kubiak M, Duda AM, Ganadu ML, Kozlowski H. Crystal structure of a copper(II)-famotidine complex and solution studies of the Cu2þ -famotidine-histidine ternary system. J Chem SocDalton Trans 1996;1905–1908.
370. Umadevi B, Muthiah PT, Shui X, Eggleston DS. Metal–drug interactions: Synthesis and crystal structure of dichlorodithiabendazolecobalt(II) monohydrate. Inorg Chim Acta 1995;234:149–152.
371. Sanchez-del Grado RA, Navarro M, Perez H, Urbina JA. Toward a novel metal-based chemotherapy against tropical diseases. 2. Synthesis and antimalarial activity in vitro and in vivo of new ruthenium- and rhodium-chloroquine complexes. J Med Chem 1996;39:1095–1099.
372. Clarke MJ, Stubbs M. Interactions of metallopharmaceuticals with DNA. Met Ion Biol Sys 1996;32:727– 373. Nacsa J, Nagy L, Molnar J, Molnar J. Trifluoperazine and its metal complexes inhibit the Moloney leukemia virus reverse transcriptase. Anticancer Res 1998;18:1373–1376.
374. Kostova IP, Changov LS, Keuleyan EE, Gergova RT, Manolov II. Synthesis, analysis and in vitro antibacterial activity of new metal complexes of sulbactam. Farmaco 1989;53:737–740.
375. Jackson A, Davis J, Pither RJ, Rodger A, Hannon MJ. Estrogen-derived steroidal metal complexes: Agents for cellular delivery of metal centers to estrogen receptor-positive cells. Inorg Chem 2001;40:3964–3973.
376. (a) Teitz Y, Barko N, Abramolf W, Ronen D. Relationships between structure and antiretroviral activity of thiosemicarbazone derivatives. Chemotherapy 1994;40:195–200. (b) Winter J, Hall RL, Moyer RW. Theeffect of inhibitors on the growth of the entomopoxvirus from Amsacta moorei in Lymantria dispar (gypsymoth) cells. Virology 1995;211:462–473.
377. Rodrı´guez-Argu¨elles MC, Sa´nchez A, Ferrari MB, Fava GG, Pelizzi C, Pelosi G, Albertini R, Lunghi P, Pinelli S. Transition-metal complexes of isatin-b-thiosemicarbazone. X-ray crystal structure of two nickelcomplexes. J Inorg Biochem 1999;73:7–15.
378. (a) Supuran CT, Scozzafava A, Saramet I, Banciu MD. Carbonic anhydrase inhibitors: Inhibition of isozymes I, II, and IV with heterocyclic mercaptans, sulfenamides, sulfonamides, and their metal com-plexes. J Enzy Inhib 1998;13:177–194. (b) Supuran CT, Scozzafava A, Mincione F, Menabuoni L, BrigantiF, Mincione G, Jitianu M. Carbonic anhydrase inhibitors. Part 60. The topical intraocular pressure-loweringproperties of metal complexes of a heterocyclic sulfonamide: Influence of the metal ion upon biologicalactivity. Eur J Med Chem 1999;34:585–595. (c) Ilies M, Supuran C, Scozzafava A. Carbonic anhydraseinhibitors. Part 91. Metal complexes of heterocyclic sulfonamides as potential pharmacological agents inthe treatment of gastric acid secretion imbalances. Metal-Based Drugs 2000;7:57–62. (d) Briganti F, TilliS, Mincione G, Mincione F, Menabuoni L, Supuran CT. Carbonic anhydrase inhibitors. Metal complexes of5-(2-chlorophenyl)-1,3,4-thiadiazole-2-sulfonamide with topical intraocular pressure lowering properties:The influence of metal ions upon the pharmacological activity. J Enzyme Inhib 2000;15:185–200.
379. Orvig C, Abrams MJ, Eds. Medicinal inorganic chemistry. Chem Rev 1999;99(9): Special issue.
380. Sadler PJ. Inorganic chemistry and drug design. Adv Inorg Chem 1991;36:1–48.
381. Birch NJ. Lithium: Inorganic pharmacology and psychiatric use. Oxford: IRL Press; 1988. (b) Birch NJ.
Inorganic pharmacology of lithium. Chem Rev 1999;99:2659–2682. (c) Lenox RH, Hahn CG. Overview ofthe mechanism of action of lithium in the brain: Fifty-year update. J Clin Psychiat 2000;61(Suppl 9):5–15.
382. Moore GJ, Bebchuk JM, Wilds IB, Chen G, Menji HK. Lithium-induced increase in human brain grey matter. Lancet 2000;356:1241–1242. (b) Sassi RB, Nicoletti M, Brambilla P, Mallinger AG, Frank E,Kupfer DJ, Keshavan MS, Soares JC. Increased gray matter volume in lithium-treated bipolar disorderpatients. Neurosci Lett 2002;329:243–245.
383. Mota de Freitas D. Alkali metal nuclear magnetic resonance. Meth Enzymol 1993;227:78–106.
384. Yan S, Ding K, Zhang L, Sun H. Complexation of antimony(III) by trypanothione. Angew Chem Int Ed Professor Li-June Ming completed his undergraduate education at the Chinese Culture University in Taiwan,earned his M.S. degree on fast kinetic study of configurational change of metal complexes with Professor Chung-Sun Chung at the National Tsing Hua University in Taiwan, and obtained his Ph.D. degree on nuclear magneticresonance studies of Cu,Zn-superoxide dismutase with Professor Joan Valentine at the University of California atLos Angeles in 1988. After 3 years of postdoctoral research on several Fe enzymes and synthetic complexes withProfessor Lawrence Que, Jr., at the University of Minnesota, he joined the Chemistry Department and theInstitute for Biomolecular Science at the University of South Florida in 1991. His research interest is onhydrolytic metalloenzymes and chemistry, metalloantibiotics, and metallopolymers. Nuclear magnetic reso-nance techniques have been extensively used in his research for the investigation of paramagneticmetallobiomolecules and synthetic metal complexes.

Source: http://chuma.cas.usf.edu/~ming/Publications/abstracts/metalloantibiotics.pdf

Microsoft word - technical specs chapt 7 pass process

PASS CR (MICRO Type II) CAPE SEAL and PASS QB Rejuvenating Seal for Residential Roads SECTION 700 - PASS CR Scrub Seal The work shall consist of furnishing all necessary labor, materials and equipment for the transporting, application of the polymer modified asphaltic emulsion PASS or equal, ¼" by No. 10 premium aggregate to conform to the Provisions of Section 37-2, of the Standard Specifications, Plans and these Special Provisions. The work shall be done in the following order: preparing the pavement surface; applying the emulsion; scrubbing the applied emulsion with an emulsion broom; applying premium aggregate; rolling the ¼" by No. 10 premium aggregate; and sweeping up excess aggregate and more fully described below.

Clasificacin terminolgica

CLASIFICACIÓN TERMINOLÓGICA Y CODIFICACIÓN DE ACTOS Y TÉCNICAS MÉDICAS ORGANIZACIÓN MÉDICA COLEGIAL CODIFICACIÓN DE ESPECIALIDADES OMC La codificación de especialidades en el marco de la OMC se realiza mediante un código numérico de 2 dígitos, empezando por Medicina General como 01 y Pediatría como 02, para seguir con las sucesivas especialidades por orden alfabético desde 03 correspondiente a Alergología hasta el 41 para Urología.