* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download full text pdf
Psychopharmacology wikipedia , lookup
Pharmacogenomics wikipedia , lookup
Discovery and development of direct Xa inhibitors wikipedia , lookup
Discovery and development of cephalosporins wikipedia , lookup
Prescription costs wikipedia , lookup
NK1 receptor antagonist wikipedia , lookup
Discovery and development of tubulin inhibitors wikipedia , lookup
Discovery and development of proton pump inhibitors wikipedia , lookup
Pharmaceutical industry wikipedia , lookup
Toxicodynamics wikipedia , lookup
Discovery and development of neuraminidase inhibitors wikipedia , lookup
Drug interaction wikipedia , lookup
DNA-encoded chemical library wikipedia , lookup
Discovery and development of non-nucleoside reverse-transcriptase inhibitors wikipedia , lookup
Neuropsychopharmacology wikipedia , lookup
Neuropharmacology wikipedia , lookup
Pharmacognosy wikipedia , lookup
Discovery and development of ACE inhibitors wikipedia , lookup
Discovery and development of antiandrogens wikipedia , lookup
Discovery and development of integrase inhibitors wikipedia , lookup
Metallodrugs 2014; 1, 10–23 Research Article Open Access Seth S. Bradford, J. A. Cowan From Traditional Drug Design to Catalytic Metallodrugs: A Brief History of the Use of Metals in Medicine Abstract: Traditional drug design has been effective in the development of therapies for a variety of disease states but there is a need for new approaches that will tackle new challenges and complement current paradigms. The use of metals in medicine has resulted in several successes and allows for the introduction of properties that cannot be achieved by use of organic compounds alone, but also introduces new challenges that can be addressed by a careful understanding of the principles of inorganic chemistry. Toward this end, the unique structural and coordination chemistry, as well as the reactivity of metals, has been used to design novel classes of therapeutic and diagnostic agents. This review briefly summarizes progress in the field of therapeutics, from the earliest use of metals to more recent efforts to design catalytic metallodrugs that promote the irreversible inactivation of therapeutically relevant targets. Keywords: drug design, enzyme inhibitors, metals in medicine, catalytic, metallodrug DOI 10.2478/medr-2014-0002 1 Introduction History is filled with descriptions of disease states and ailments, as well as attempts to combat them. The use of chemical substances for the treatment of these conditions *Corresponding author: J. A. Cowan: Evans Laboratory of Chemistry, The Ohio State University, 100 West 18th Avenue, Columbus, Ohio 43210. Tel: 614 292 2703; Fax: 614 292 1685; e-mail: cowan@ chemistry.ohio-state.edu Seth S. Bradford: Evans Laboratory of Chemistry, The Ohio State University, 100 West 18th Avenue, Columbus, Ohio 43210 J. A. Cowan: MetalloPharm LLC, 1790 Riverstone Drive, Delaware, OH 43015 most likely stemmed from observations of cause and effect, such as the discovery that a plant had healing properties, or that it relieved pain. For example, the use of salicylic acid-rich plants for pain and inflammation dates back to before the Common Era, but it wasn’t until the mid to late 1800’s that the active component was identified as acetylsalicylic acid and marketed as aspirin [1]. More recent advances in the sciences have allowed for the determination of the exact chemical composition of drugs as well as an appreciation of the importance of the drug target. Essential techniques for studying the effects of a drug on its target, both in the test tube and in vivo, required key advances in basic science, such as the development of the microscope and discovery of various microorganisms, including bacteria by Antony Leeuwenhoek in the 1600’s [2], the discovery of cells in 1665 by Robert Hooke [3], the development of staining procedures, such as the Gram stain, for cells and tissues developed in the 1700’s and 1800’s [4], and the identification of viruses by Dmitri Ivanovsky and Martinus Beijerinck in the late 1800’s [5]. Treatment also required a deeper understanding of the origins of the disease. Malaria has been known for over 4,000 years, but the malaria parasite wasn’t identified until 1880 by Charles Louis Alphonse Laveran [6]. As the molecular understanding of drugs and drug targets has improved, so have the methodologies for optimizing and testing them. Current approaches to the discovery of natural products rely on the screening, separation, characterization, and synthesis of the specific molecules involved in achieving the desired response. The use of drugs derived from nature remains an important aspect of pharmaceutical development, but advances in the sciences have also allowed for an improved understanding of how drugs work and how best to administer them. The rational design of drugs derives from a basic understanding of the cause of a condition, as well as the ability to alter the processes that ultimately led to it. As advances in chemistry and the biological sciences have continued, and a deeper understanding of drug © 2014 Seth S. Bradford, J. A. Cowan, licensee De Gruyter Open. This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivs 3.0 License. Unauthenticated Download Date | 6/15/17 8:02 PM Past, Current and Future Use of Metals in Medicine interactions with a target has developed, the rational design of drugs becomes more plausible. A compound can be modified to alter its properties to generate higher activity, lower toxicity, improved specificity, better stability, or an improved half-life in vivo. Modifications can be made in a systematic way to generate structure activity relationships (SARs) that help to guide further optimization [7]. One of the earliest examples of a medicinal chemistry approach to drug design is illustrated by the work of Paul Ehrlich on studies of arsenic compounds for the treatment of syphilis [8]. Ehrlich’s concept of a “magic bullet” that selectively targets a bacterium over other organisms is an important concept even today, and achieving this selectivity is still one of the most promising, yet challenging, aspects of drug design. 2 Traditional Drug Design Currently, the development of drugs usually involves a combination of screening to identify lead compounds, as well as rational design to further optimize these compounds. Rational design acts as a guiding hand to focus research, but the complexities ultimately make prediction an uncertain process and there is still a great deal of trial and error involved. A pharmaceutical company will often have access to a very large number of compounds that need to be narrowed down according to other parameters such as computational results or physical properties. Lipinski’s rule of five is commonly used to determine the “druglikeness” of molecules [9]. It 11 does not matter how active a compound is in an assay if it is insoluble or has other properties that would preclude it from being used in vivo. Lipinski found that most of the medications in use consisted of compounds with no more than five hydrogen bond donors and 10 hydrogen bond acceptors, had a molecular weight less than 500 grams per mole, and had an octanol-water partition coefficient (log P) of less than five. These rules are used as a guideline to focus the screening process and new drugs tend to be small organic molecules with high affinity for their target as well as good stability and solubility in vivo. It is important to keep in mind, however, that these are only guidelines. The recently approved protease inhibitor for the treatment of hepatitis C, telaprevir, is a notable example of a compound that breaks all of Lipinski’s rules. The most common drug targets are enzymes. For example, a common approach to blood pressure regulation is inhibition of angiotensin converting enzyme (ACE). ACE promotes an increase in blood pressure by converting angiotensin I to angiotensin II, a vasoconstrictor, and also by degrading the vasodilator bradykinin [10]. Proper regulation of this enzyme is key to good health. In situations where blood pressure is high, inhibition of ACE by small molecules is a common treatment. This is how some commonly used blood pressure medications such as captopril (Capoten) [11] and lisinopril (Zestril, Tensopril) (Figure 1) [12] work. The general scheme for this type of inhibition is shown in Figure 2 and is the most common approach to drug design. The small molecule inhibitor (I) binds to the active site of the enzyme (E) and blocks it so that the substrate (S) cannot bind. H 2N HS HO O O O H H N N N H OH O OH O Lisinopril Captopril Figure 1: Known small molecule inhibitors of angiotensin converting enzyme. E+S + ES + I I E+P K'i Ki EI ESI Figure 2: Reversible inhibition of an enzyme (E) that acts on a substrate (S) by an inhibitor (I). Unauthenticated Download Date | 6/15/17 8:02 PM 12 S.S. Bradford, J.A. Cowan There are many advantages to traditional drug design and the majority of the drugs currently on the market were discovered and developed in this way. The scheme shown in Figure 2 shows a series of equilibrating reactions and the requirement for a sufficient amount of inhibitor (I) to shift the equilibrium and achieve the desired effect. This can lead to a need for higher concentrations of drug, potentially more frequently, leading to toxicity and issues with patient compliance. Also, because it is reversible, the inhibitor will be released and be cleared by the body, leaving the enzyme to start converting substrate to product all over again. The use of a small number of elements in current drugs, primarily C, H, N, O, and S, limits the range of available properties available to them. The shapes that can be achieved, for example to fit into an enzyme active site, will be limited to geometries available to those atoms whereas higher coordination numbers, and hence more diverse shapes, can be attained with compounds that take advantage of a wider range of the periodic table. The use of metals in medicine is a less traditional approach which addresses some of these limitations while introducing disadvantages of its own, such as considerations of the stability and toxicity of the complex. Proper ligand design and application of basic inorganic principles, however, allows for the design of stable compounds with a wide array of properties available to them. The focus of this review, following a historical perspective, is on actual, or potential, metal-based drugs that promote therapeutic activity through direct metal-centered binding or reactivity. Other approaches that employ outer sphere contacts of coordination complexes or reactivity toward ill-defined cellular targets are not included for reasons of space limitations, although interested readers are directed to other informative reviews and articles [13-20]. 3 Metals in Medicine The importance of metals in biology has become apparent over time as an increasing number of enzymes and O HS O other biological molecules have been identified that contain metals that are important, if not essential, for structure and/or function. Metals can perform functions that cannot otherwise be achieved, such as electrolyte balance (important for processes that include the firing of neurons), as well as oxygen transport, electron transfer, and catalysis. The importance of metals to health is made clear by the abundance of diseases caused by an excess or deficiency of a metal. For example, iron is essential for life, but in humans too much iron can lead to haemochromatosis and too little iron can lead to iron deficiency. Therefore, iron regulation is essential. Other examples of diseases related to metal ion regulation include Wilson’s disease [21], Friedrich’s ataxia [22], and hypozincemia [23]. Metal overload can often be treated by use of metal chelating compounds, and British antiLewisite (dimercaprol, Figure 3) was the first chelating compound used medicinally [24]. Metal overload can also be caused by intake of metals either involuntarily or voluntarily, as was often the case in historical examples of the use of metals to treat disease. Silver has found many uses in medicine, most notably as an antibiotic, although its mechanism of action remains unclear. Prior to the development of more effective and targeted antibiotics around the time of World War I, colloidal silver was used as a disinfectant and it still finds use today in the form of silver sulfadiazine (Figure 3) for the treatment of burn wounds [25]. Organotin compounds find use as pesticides but their use has been limited by toxicity concerns [26]. The use of mercury is another common example of a compound that was used in the past without a full understanding of how it works [27]. Probably the most commonly known use of mercury was in the processing of felt hats which caused toxicity and led to the term “mad as a hatter” [28]. Medicinally, it was administered in the form of mercury (I) chloride as a disinfectant and diuretic and later found use in the form of mercury (II) chloride for the treatment of syphilis. These compounds showed some activity but their major downfall was that they were administered without a full understanding of their N S N OH HS Ag H 2N Dimercaprol N + Silver Sulfadiazine Figure 3: Metal related therapies used throughout history. Unauthenticated Download Date | 6/15/17 8:02 PM toxicity. They also illustrate the potential for activity from inorganic compounds, but also the possible pitfalls when used without taking proper precautions and without a clear understanding of their mechanism of action. While there are some metal-based therapeutics currently used, they often tend to be the exception rather than the rule and tend to be discovered rather than rationally designed. The potential for toxicity from the metal center is a serious concern, but one that can be addressed by proper ligand design and knowledge of coordination chemistry. Nevertheless, it is important that the use of a metal center is necessary for the chosen application, such as for the introduction of properties that cannot be achieved by a traditional organic compound. Examples include the use of metals for imaging, such as a positron emitting metal for positron emission tomography (PET) [29], a gadolinium complex for MRI contrast [30], or for the introduction of new reactivity such as in bleomycin [31]. Another potential use of metals is to generate new shapes that can more effectively explore structure activity space [32], such as the use of an inert metal center to introduce an octahedral geometry that better fits the active site of protein kinases. In the current market, former problems and perceptions associated with metal toxicity are less likely to be a major issue, due to an increased emphasis on proper testing and understanding of the Past, Current and Future Use of Metals in Medicine 13 mechanism of potential therapeutics. There are several examples of metal-based drugs that are still currently in use and these compounds illustrate both the potential of this class of compounds and the need for further work to optimize their uses in a clinical setting. 3.1 Overview of Metal-Based Therapeutics in Current Use The number of metal-based therapeutics that have received FDA-approval is limited and this is a reflection both of the challenges involved and the lack of interest of pharmaceutical companies in inorganic compounds. There is precedent, however, for a willingness to approve a compound quickly, especially if it fits an identified need that is not being met. A good example of this is the large number of imaging agents currently on the market, such as the imaging agent Prohance (Figure 4) which was approved in 1992 and is an MRI contrast agent used for imaging of the central nervous system [33]. Gadolinium contains seven unpaired electrons and interacts with water to enhance its relaxivity. This enhancement occurs only where the gadolinium is present and can be used to image where the complex localizes in the body. The history of FDA approved metal compounds used for therapy is more limited. The use of lithium carbonate for Figure 4: Clinically used inorganic based drugs. Unauthenticated Download Date | 6/15/17 8:02 PM 14 S.S. Bradford, J.A. Cowan drugs shows the potential for generating compounds that have unique activity, but also emphasizes the limitations and potential challenges involved in controlling the toxicity and establishing a detailed understanding of the mechanism of action. the treatment of depression, primarily bipolar disorder, was approved in 1970 [34]. Its mechanism of action is still not well understood. In 1985, the gold compound auranofin (Ridaura, Figure 4) was approved for the treatment of rheumatoid arthritis. Its primary mechanism of action is thought to be inhibition of an esterase that breaks down cartilage tissue but this mechanism of action is still debated [35]. Bismuth salicylate (Figure 4), the active ingredient in Pepto-Bismol, was approved in 1996 for use as an antacid and antidiarrheal [36]. Bismuth has also found use as a treatment for leishmaniasis in the form of meglumine antimoniate (Glucantim) [37]. Another inorganic compound, sodium stibogluconate (Pentostam), has also been used in the treatment of leishmaniasis and is still used worldwide [38]. In 1997, Sclerosol Intrapleural Aerosol, Mg3Si4O10(OH)2, was approved for treatment of malignant pleural effusion in lung cancers [39]. Malignant pleural effusion is a condition where an abnormal amount of fluid collects in the lungs, and this compound is thought to induce an inflammatory response and to act as a sealant to prevent the further accumulation of liquids. Another important field of research in the application of metals to medicine is photodynamic therapy, exemplified by Photofrin, approved in 1995 as a photosensitizing agent for the treatment of esophageal cancer and later for the treatment of non-small cell lung cancer [40]. Trisenox is a formulation of As4O6 that was approved in 2000 for administration in combination with all-trans retinoic acid (ATRA) for the treatment of some forms of leukemia [41]. The metal-based radiopharmaceutical Zevalin (Figure 4) was approved in 2002 for the treatment of low-grade, B-cell, non-Hodgkin’s lymphoma [42]. This compound consists of the monoclonal mouse IgG1 antibody ibritumomab linked to the metal binding domain tiuxetan, which coordinates either indium-111 or yttrium-90. Zevalin binds to the CD20 antigen found on the surface of B cells and selectively kills them. Other therapeutic radiopharmaceuticals include Metastron, a 89Sr complex used for skeletal metastases, Quadramet, a 153Sm complex also used for osteoblastic bone metastases, and Xofigo, a 223Ra complex used for castration-resistant prostate cancer [43-45]. This range of H 3N Cl Probably the most successful metal-based drugs are represented by cisplatin and its analogs. Like many of the drugs mentioned, and many drugs in general, cisplatin (structure shown in Figure 5) was not rationally designed but serendipitously discovered by Barnett Rosenberg in 1965 [46]. Cisplatin was approved by the FDA in 1978 for the treatment of testicular and ovarian cancers. It is delivered as a neutral compound, but is susceptible to hydrolysis as shown in Figure 5. In the blood, the chloride concentration is sufficiently high that the equilibrium lies to the neutral compound on the left. Upon entering the cell, the [Cl-] drops and the cisplatin hydrolyzes to form the charged complex on the right which then cannot cross the cell membrane and therefore is trapped inside the cell. This is the active form of cisplatin. The detailed mechanisms of uptake and transport within the cell, as well as export pathways that provide potential resistance mechanisms, are only now becoming understood, with evidence that Pt-based drugs, in addition to passive diffusion mechanisms, can also be bound and mobilized by copper trafficking and transport proteins [47], and so the conversion to the aquated complex may involve sulfurbound protein forms as intermediates. Nevertheless, in one of the major therapeutic pathways, the active hydrated Pt salt reacts with DNA to generate inter- and intrastrand crosslinks that inhibit transcription and DNA repair, leading to apoptosis [48]. Understanding the factors that contribute to efficient targeting of DNA is an important problem with much activity [49]. There is also an increasing body of evidence that suggests that alternative cellular targets may be important for therapeutic activity of platinum compounds [50,51]. Understanding the chemistry of platinum drugs with cellular and serum H 3N + 2 H 2O Pt H 3N 0 3.2 Cisplatin Cl OH + 2 Cl - + Pt H 3N + H+ OH 2 cisplatin Figure 5: Hydrolysis of cisplatin to generate the active form. Unauthenticated Download Date | 6/15/17 8:02 PM Past, Current and Future Use of Metals in Medicine proteins, and the impact on cellular toxicology and pharmacology, is an area of developing interest in light of the availability of modern analytical tools [50-52]. The selectivity of cisplatin derives from an enhanced uptake in cancer cells relative to healthy cells as well as the higher accessibility of the DNA in cells that are undergoing division. This is why other cells that are continuously dividing, such as hair cells or cells in the gastrointestinal tract, are also damaged by cisplatin and this is the basis for common side effects such as hair loss and nausea. Therefore, while cisplatin does have a preference for killing cancer cells, it does not have a high level of specificity. Many derivatives of cisplatin have been studied and two have received approval by the FDA. Carboplatin (Figure 6) was approved in 1989 for the treatment of ovarian cancer and oxaliplatin [53] was approved in 2002 for the treatment of advanced colorectal cancer. Both compounds replace the chloride leaving group of cisplatin with a bidentate dicarboxylate ligand that is less reactive. This results in fewer side reactions, and therefore lower toxicity, particularly with proteins, and also better retention in the body relative to cisplatin. Cisplatin and its derivatives are a good example of the promise of inorganic pharmaceuticals and they also illustrate the concerns that need to be addressed by future metal based drugs. Their effectiveness and contributions to human health are undeniable but there is room for improvement in specificity in order to achieve lower toxicity. Resistance to cisplatin is also a major hurdle to overcome and novel approaches to treating cisplatin-resistance are needed, such as the use of gold(III) porphyrin complexes [17], as well as iridium and ruthenium compounds [16,20]. Multiple side reactions for platinum complexes, and the fact that the drug is irreversibly consumed by its mode of action, lead to a requirement for high doses, and targeted approaches are also being explored [14,19]. Achieving selective targeting to the cells of interest is still one of the biggest challenges in the use of metals in medicine as well as in drug design in general. 15 3.3 Bleomycin Bleomycin (Figure 7) was isolated from Streptomyces verticillus in 1966 and is an example of a drug designed by nature that takes advantage of the reactivity of metals [54]. It received approval in the United States in 1973 and is used for the treatment of squamous cell carcinomas, testicular cancer, and Hodgkin’s lymphoma. Bleomycin is not actually administered as a metal complex but instead consists of several domains, including one that is important for binding to DNA and another that coordinates to metals. The mechanism of action of bleomycin is the induction of DNA strand breaks due to recruitment of a metal, probably iron, and subsequent generation of reactive oxygen species which react with the DNA [55]. Much like cisplatin, bleomycin shows a preference for cancer cells but its mechanism of action is not specific to them. It is a good example, though, of a drug incorporating a metal ion to promote chemistry on a target. It is also another example of a metal-based drug that would benefit greatly from enhanced selectivity for cancer cells. 4 Catalytic Metallodrugs Cisplatin and bleomycin are two examples of toxic compounds that show some level of selectivity for a disease state but would benefit greatly from more specific targeting to the site of interest. As an extension of early work on artificial proteases and nucleases [5667], catalytic metallodrugs were developed as a class of metal-based drugs that helped to address some of these concerns [68]. The basic design strategy is shown in Figure 8 and consists of a metal binding domain linked to a target recognition domain. The target recognition domain binds to the therapeutic target of interest and provides selectivity while the metal binding domain performs chemistry on that target. This simple modular design allows for the variation of the different domains O H 3N H2 N O Pt H 3N O O Pt N H2 O O carboplatin O O oxaliplatin Figure 6: Clinically approved derivatives of cisplatin. Unauthenticated Download Date | 6/15/17 8:02 PM 16 S.S. Bradford, J.A. Cowan Figure 7: Structure of bleomycin identifying domains primarily involved in association with DNA and metal coordination (the imidazole ring has also been implicated in metal binding). to modify the properties of the metallodrug and can be quickly adapted to a variety of different therapeutic targets. As shown in the reaction scheme at the bottom of Figure 8, the metal-ligand complex (ML) binds to its target (T) in a process that is under equilibrium and analogous to the enzyme inhibition strategy described above. The next step, however, involves irreversible reaction with the target to generate the product (P). In this way it acts like an enzyme. This irreversible step pulls the equilibrium to the right, according to Le’Chatelier’s principle, and aids the progress of the reaction even under situations where binding might not be as tight as would be required for a traditional drug. The final step involves release of the metal-ligand complex from the target to regenerate the metallodrug. Hence, another important design feature is the ability to act catalytically. One of the major advantages of this approach is its ability to address some of the issues previously mentioned in terms of specificity and toxicity. One feature that emerges from the design as described is a double filter mechanism for reactivity, as shown in Figure 9. Drugs will often have side effects and toxicity due to nonspecific interactions with macromolecules other than their therapeutic target. For catalytic metallodrugs, however, there is a requirement for both binding and proper positioning of the metal binding site for reactivity. Even if nonspecific binding occurs, as shown on the right, if the metal binding domain isn’t properly oriented to react with the target then it will not be active. This extra level of selectivity addresses one of the biggest issues with the currently used metal-based drugs described above, which is their inherent toxicity. This feature, combined with the fact that these metallodrugs are designed to act catalytically, allows for higher selectivity at a lower dose. Lower dosages should also contribute to lower costs and, combined with reduced toxicity, better patient compliance. The unique mechanism of action also makes catalytic metallodrugs a perfect complement to current therapies both in terms of combination therapy and to combat resistance. The novel mechanism of action makes it very difficult for resistance to develop against both a catalytic metallodrug and another drug at the same time. Often resistance can occur by mutation of the therapeutic target to prevent or reduce binding. In the case of catalytic metallodrugs, however, this reduction in binding could potentially not affect, or even enhance, catalyst activity depending on what the rate limiting step is for a particular system. 4.1 The Metal Binding Domain As mentioned, the role of the metal binding motif is to perform chemistry on the therapeutic target. In order to be useful as a metallodrug, it must also bind to the Unauthenticated Download Date | 6/15/17 8:02 PM Past, Current and Future Use of Metals in Medicine 17 Figure 8: (Top) Basic catalytic metallodrug design. (Bottom) Reaction scheme for catalytic inactivation of target T by the catalytic metallodrug consisting of a metal binding domain (M) and a target recognition domain (L). Figure 9: Double filter mechanism for catalytic metallodrugs [89]. metal very tightly and be stable both kinetically and thermodynamically. This domain allows for further customization and optimization of the metallodrug. Selection criteria include factors such as intrinsic reactivity, stability, and compatibility with the targeting domain. Selection of the metal binding domain will also determine the range of metals that can be used another factor that can be varied to optimize therapeutic activity. Other common metal binding ligands include ethylenediaminetetraacetic acid (EDTA), terpyridine, cyclen and its derivatives, as well as porphyrins and expanded porphyrins. Examples of the use of these metal binding domains are discussed later in section 4.3. Another metal binding domain that has been used extensively in design efforts is the amino terminal copper and nickel binding (ATCUN) motif derived from the N-terminus of serum albumins (Figure 10) [69]. It has been shown to bind to copper and nickel very tightly (KD for copper of 1.18 x 10-16 M) [70], with precedence for reaction with nucleic acids [71]. Copper and nickel coordination is promoted by binding of the N-terminal amine, two backbone amides, and the histidine side chain in the third position (Figure 10). NMR, EPR, visible spectroscopy and X-ray crystallography are consistent with a divalent metal ion coordinated in a square planar configuration [70,72-74], leaving two faces accessible for binding to a Unauthenticated Download Date | 6/15/17 8:02 PM 18 S.S. Bradford, J.A. Cowan Figure 10: The amino terminal copper and nickel binding (ATCUN) motif. target molecule. This design allows for variation of the properties of this domain by changing the side chains (R and R’, Figure 10) contributed by the first two residues. Both copper and nickel undergo redox chemistry through the M3+/2+ couple rather than M2+/+ [75] lending the complex greater reactivity and thermodynamic stability. Reaction pathways appear to vary from system to system, but usually involve the generation of reactive oxygen species through pathways such as those shown in Figure 11, and can involve superoxide, peroxide, hydroxyl radicals, or metal bound oxygen species. Coreagents can also be important for cycling and regenerating the metal complex in the starting state, such as by reduction of Cu3+ to Cu2+ by ascorbic acid, or can be important for other related redox processes. These conditions are expected to be present in some form under physiological conditions since ascorbic acid (as vitamin C), hydrogen peroxide, and many other reducing and oxidizing agents are available in vivo. 4.2 The Target Recognition Domain and Linker The target recognition domain is anything that can bind to and recognize a specific target of interest. While tight binding is generally desirable, this could inhibit turnover through product inhibition and lead to a less effective catalyst if a high affinity resulted in a slow off-rate for catalyst release. More important than tight binding, however, is specificity, which can be retained even with weaker binding through a property described as the double filter effect (Figure 9). This recognizes the fact that in the case of a catalytic drug that promotes chemical change on the target molecule, specificity will reflect not only Figure 11: Relevant reduction potentials (relative to NHE) for oxidative chemistry of catalytic metallodrugs. Unauthenticated Download Date | 6/15/17 8:02 PM Past, Current and Future Use of Metals in Medicine binding propensity, but also an appropriate orientation for efficient chemistry to occur. Such a combination is less likely to arise for off-target binding, and so specificity is maintained. The design of metallodrugs is facilitated by possession of functionality on the targeting molecule that can be used to couple with a metal binding domain, but without significantly perturbing recognition and binding to the therapeutic target. This would most likely be a small molecule that is already known to bind to a particular target and would benefit from the reactivity introduced by the metal binding domain, but other larger biological molecules, such as proteins or peptides, including antibodies, nucleic acids, or carbohydrates are not excluded. Targeting domains that are based on peptides can readily incorporate the ATCUN motif described above, but any compound that can be modified to allow for conjugation to the metal binding domain can be elaborated. The design can also include a linker moiety that separates the metal binding and targeting domains and can be used to further tailor reactivity. For example, the introduction of flexibility could either enhance or reduce chemistry at the target by influencing the placement and orientation of the reactive metal center relative to potential scissile bonds. This linker can also be varied to tailor other properties, such as water solubility, partition coefficient, or in vivo half-life. Depending on the mode of binding, other Trp123 12 A functional groups can be incorporated to further tailor the properties of the catalytic metallodrug, including the optimization of traditional in vivo and drug-like properties such as the pharmacokinetic profile, cellular uptake, and oral availability. 4.3 Development of Catalytic Metallodrugs Among the first catalytic metallodrugs reported were compounds acting as inhibitors of a class of metalloproteases involved in regulation of the cardiovascular system, such as angiotensin-converting enzyme (ACE) and endothelin-converting enzyme 1 (ECE-1) [76]. These compounds, however, acted by this new mechanism of catalytic inactivation rather than reversible inhibition alone. Further studies with the model enzyme carbonic anhydrase I (CA-I) demonstrated that copperATCUN complexes of the known inhibitor sulfanilamide caused a time-dependent inactivation of CA-I that was further shown by enzyme digestion and LC/MS/MS analysis to be due to irreversible modification of active site residues important for the enzyme mechanism (Figure 12, left) [77]. More recent ACE targeted metallodrugs based on the known inhibitor lisinopril (Figure 12, right) showed both concentration- and time-dependent inactivation of ACE [78]. These reactions occur by an oxidative mechanism which, although the details vary from system to system, involve the generation of reactive oxygen species that His103 27 A His243 19 A His64 8 A His67 6 A His200 5.5 A 19 Trp97 15 A Zn2+ His40 31 A Figure 12: (left) Residues in carbonic anhydrase that are modified by Cu-GGH conjugates of sulfanilamide. His64, His67, and His200 have been implicated in rate-limiting proton transfer within the active site [77]. Conjugates of lisinopril (right, where R = lisinopril) with a metal binding domain have been shown to catalytically inactivate angiotensin converting enzyme [78]. Unauthenticated Download Date | 6/15/17 8:02 PM S.S. Bradford, J.A. Cowan species. Further work has demonstrated the targeted cleavage of RNA critical to HIV-1 replication, the Rev Response Element (RRE), and provided a series of compounds with a unique mechanism of action that could find potential use in the treatment of HIV [85,86]. Compounds based on Cu-GGH-YrFK-amide were shown to target stem loop IIb of the Hepatitis C Virus (HCV) internal ribosomal entry site (IRES), which is important for viral translation, and catalytically degrade it [87]. These compounds demonstrated high activity both in vitro and in a cellular replicon assay that, in the absence of a proven animal model for HCV, is accepted by the USFDA as a measure of drug efficacy (Figure 13, left). This compound showed additive, possibly slightly synergic activity, when given in combination with recombinant interferon α-2b (rIFNα-2b). Site specific oxidative degradation of the target RNA was demonstrated and the most likely mechanism of action for these compounds is a mixed oxidative/ hydrolytic mechanism where generation of the copper (III) center is necessary to enhance the Lewis acidity of the metal ion. Real-time polymerase chain reaction (RTPCR) experiments showed that the mechanism in cells is consistent with the proposed mechanism of HCV RNA cleavage (Figure 13, right) [88]. 5 Conclusions Catalytic metallodrugs fill a current need in drug design by introducing new properties and providing a new class of drugs that act by a unique mechanism of action. They also 1.2 1.0 % of Virus Control (Efficacy) Normalized Copies of HCV RNA ultimately inactivate the target of interest. Research in the Suh lab has taken a different approach to the ACE system, designing Co(III) catalysts that target the angiotensin I and I peptides rather than the enzyme itself [79], a reaction that is likely to occur by a hydrolytic mechanism. Similarly, another cobalt complex was identified that could cleave both human islet amyloid polypeptide, important in type 2 diabetes mellitus, and soluble oligomers of amyloidogenic peptides amyloid β-42, thought to be involved in the pathogenesis of Alzheimer’s disease [80]. Work in the Blum lab also demonstrated the use of copper and zinc compounds for the irreversible inhibition of furin, an enzyme that has been implicated in a variety of disease states including HIV, anthrax infection, and cancer [81]. Catalytic metallodrugs targeting nucleic acids have also been reported. Nucleic acids are more susceptible to hydrolysis and much of the work in this area has focused on artificial nucleases that cleave the phosphodiester bond, as exemplified by work from Morrow and Richards and coworkers [62]. There is still precedence, however for the oxidative cleavage of nucleic acids and, in fact, many of the catalytic metallodrugs that incorporate the ATCUN motif react via oxidation of DNA [59]; this nuclease activity has been implicated in the mechanism of antimicrobial peptides [82]. The texaphyrins pioneered in the Sessler lab also act through reactive oxygen species (ROS) that react with DNA and lead to cell death [83]. Another unique approach to the site specific targeting of DNA is demonstrated by the work of Komiyama et al. [84] wherein the target recognition domain is a peptide nucleic acid and cleavage is performed by a catalytic CeIV/EDTA 0.8 0.6 0.4 0.2 0.0 0 3 6 Day 9 0.2 ΙU/mL rIFNµ-2b 110 100 0.2 µM Complex 90 80 70 60 50 40 30 Both 20 10 0 I II III % Virus Control Drugs Evaluated 110 100 90 80 70 60 50 40 30 20 10 0 % of Cell Control (Toxicity) 20 % Cell Control Figure 13: (left) RT-PCR results showing concentration and time-dependent degradation of HCV RNA in the replicon assays. 0 μM (black square), 2.5 μM (white circle), 5 μM (black triangle), 10 μM (white diamond), 20 μM (black star). The graph is normalized to the RNA levels in the absence of metallopeptide [87]. (right) HCV cellular replicon results showing additive to slightly synergic activity of the metallodrug CuGGHYrFK-amide when given in combination with pegylated interferon. (I, 0.2 μM 1, II, 0.2 IU/mL rIFNα-2b, III, 0.2 μM 1 + 0.2 IU/mL rIFNα-2b). Unauthenticated Download Date | 6/15/17 8:02 PM address some of the current drawbacks with metal-based drugs, such as a requirement for high doses and poor selectivity leading to high toxicity. The modular design also allows for the variation of metal binding and targeting components to optimize the properties of the drug. The targeting domain represents a way to achieve enhanced selectivity and also makes it generally applicable to any system of interest. Acknowledgements: This work was supported by grants from the National Institutes of Health [HL093446 and AA016712]. Seth Bradford was supported by an NIH Chemistry/Biology Interface training grant (T32 GM08512). References [1] Ibsim. 110 years history of aspirin: effectiveness and activity mechanism. Hwahak Sekye 2009, 49, 48-52. [2] Corliss J., O. A salute to Antony van Leeuwenhoek of Delft, most versatile 17th century founding father of protistology. Protist 2002, 153, 177-190. [3] Rao, K. K. Life and work of Robert Hooke (1635-1703). J. Inst. Eng. (India), Part GE 1967, 48, 42-50. [4] Popescu, A.; Doyle, R. J. The Gram stain after more than a century. Biotech. Histochem. 1996, 71, 145-151. [5] Artenstein A., W. The discovery of viruses: advancing science and medicine by challenging dogma. Int. J. Infect. Dis. 2012, 16, e470-473. [6] Bruce-Chwatt, L. J. Alphonse Laveran’s discovery 100 years ago and today’s global fight against malaria. J. R. Soc. Med. 1981, 74, 531-536. [7] Guo, Z. Structure-activity relationships in medicinal chemistry: development of drug candidates from lead compounds. Pharmacochem. Libr. 1995, 23, 299-320. [8] Thorburn, A. L. Paul Ehrlich: pioneer of chemotherapy and cure by arsenic (1854-1915). Br. J. Vener. Dis. 1983, 59, 404-405. [9] Lipinski, C. A.; Lombardo, F.; Dominy, B. W.; Feeney, P. J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Delivery Rev. 2001, 46, 3-26. [10] Izzo, J. L., Jr.; Weir, M. R. Angiotensin-converting enzyme inhibitors. J. Clin. Hypertens. 2011, 13, 667-675. [11] Opie, L. H.; Kowolik, H. The discovery of captopril: from large animals to small molecules. Cardiovasc. Res. 1995, 30, 18-25. [12] Noble, T. A.; Murray, K. M. Lisinopril: a nonsulfhydryl angiotensin-converting enzyme inhibitor. Clin. Pharm. 1988, 7, 659-669. [13] Butler, J. S.; Sadler, P. J. Targeted delivery of platinum-based anticancer complexes Curr. Opin. Chem. Biol. 2013, 17, 175-188. [14] Dhara, S.; Kolishettib, N.; J., L. S.; Farokhzadb, O. C. Targeted delivery of a cisplatin prodrug for safer and more effective prostate cancer therapy in vivo. Proc. Natl Acad. Sci. USA 2011, 108, 1850–1855. [15] Dorr, M.; Meggers, E. Metal complexes as structural templates for targeting proteins. Curr. Opin. Chem. Biol. 2014, 19, 76-81. Past, Current and Future Use of Metals in Medicine 21 [16] Hearn, J. M.; Romero-Canelón, I.; Qamar, B.; Liu, Z.; HandsPortman, I.; Sadler, P. J. Organometallic iridium(III) anticancer complexes with new mechanisms of action: NCI-60 screening, mitochondrial targeting, and apoptosis. Chem. Biol. 2013, 8, 1335−1343. [17] Lum, C. T.; Sun, R. W.-Y.; Zou, T.; Che, C.-M. Gold(III) complexes inhibit growth of cisplatin resistant ovarian cancer in association with upregulation of proapoptotic PMS2 gene. Chem. Sci. 2014, 5, 1579-1584. [18] Mjos, K. D.; Orvig, C. Metallodrugs in medicinal inorganic chemistry. Chem. Rev. 2014, 114, 4540−4545. [19] Valencia, P. M.; Pridgen, E. M.; Perea, B.; Gadde, S.; Sweeney, C.; Kantoff, P. W.; Bander, N. H.; Lippard, S. J.; Langer, R.; Karnik, R. Synergistic cytotoxicity of irinotecan and cisplatin in dual-drug targeted polymeric nanoparticles. Nanomedicine 2013, 8, 687-698. [20] Wu, K.; Liu, S.; Luo, Q.; Hu, W.; Li, X.; Wang, F.; Zheng, R.; Cui, J.; Sadler, P. J.; Xiang, J. Thymines in single-stranded pligonucleotides and G-quadruplex DNA are competitive with guanines for binding to an organoruthenium anticancer complex. Inorg. Chem. 2013, 52, 11332−11342. [21] Lorincz, M. T. Neurologic Wilson’s disease. Ann. N. Y. Acad. Sci. 2010, 1184, 173-187. [22] Anderson, G. J. Ironing out disease: inherited disorders of iron homeostasis. IUBMB Life 2001, 51, 11-17. [23] Tubek, S.; Grzanka, P.; Tubek, I. Role of zinc in hemostasis: A review. Biol. Trace Elem. Res. 2008, 121, 1-8. [24] Nash Robert, A. Metals in medicine. Altern. Ther. Health Med. 2005, 11, 18-25. [25] Nangia, A. K.; Hung, C. T.; Lim, J. K. C. Silver sulfadiazine in the management of burns - an update. Med. Actual. 1987, 21-30. [26] de Carvalho Oliveira; Santelli, R. E. Occurrence and chemical speciation analysis of organotin compounds in the environment: A review. Talanta 2010, 82, 9–24. [27] Abromowitz, E. W. Historical points of interest on the mode of action and ill effects of mercury. Bull. N. Y. Acad. Med. 1934, 10, 695-705. [28] Waldron, H. A. Did the mad hatter have mercury poisoning? Br. Med. J. (Clin. Res. Ed.) 1983, 287, 1961. [29] Li, Z.; Conti, P. S. Radiopharmaceutical chemistry for positron emission tomography. Adv. Drug Deliv. Rev. 2010, 62, 1031-1051. [30] Waters, E. A.; Wickline, S. A. Contrast agents for MRI. Basic Res. Cardiol. 2008, 103, 114-121. [31] Galm, U.; Hager, M. H.; Van Lanen, S. G.; Ju, J.; Thorson, J. S.; Shen, B. Antitumor antibiotics: bleomycin, enediynes, and mitomycin. Chem. Rev. 2005, 105, 739-758. [32] Feng, L.; Geisselbrecht, Y.; Blanck, S.; Wilbuer, A.; AtillaGokcumen, G. E.; Filippakopoulos, P.; Kraling, K.; Celik, M. A.; Harms, K.; Maksimoska, J. Structurally sophisticated octahedral metal complexes as highly selective protein kinase inhibitors. J. Am. Chem. Soc. 2011, 133, 5976-5986. [33] Kumar, K.; Tweedle, M.; Brittain, H. G. Gadoteridol. Anal. Profiles Drug Subst. Excipients 1996, 24, 209-241. [34] Pachet, A. K.; Wisniewski, A. M. The effects of lithium on cognition: an updated review. Psychopharmacology 2003, 170, 225-234. [35] Fricker, S. P. Medical uses of gold compounds: Past, present and future. Gold Bull. 1996, 29, 53-60. Unauthenticated Download Date | 6/15/17 8:02 PM 22 S.S. Bradford, J.A. Cowan [36] Bierer, D. W. Bismuth subsalicylate: history, chemistry, and safety. Rev. Infect. Dis. 1990, 12, S3-S8. [37] Sundar, S.; Chakravarty, J. Leishmaniasis: an update of current pharmacotherapy. Expert Opin. Pharmacother. 2013, 14, 53-63. [38] Murray, H. W. Leishmaniasis in the United States: treatment in 2012. Am. J. Trop. Med. Hyg. 2012, 86, 434–440. [39] Kennedy, L.; Sahn, S. A. Talc pleurodesis for the treatment of pneumothorax and pleural effusion. Chest 1994, 106, 1215-1222. [40] Moghissia, K.; Dixon, K.; Honsa, B. A.; Stringer, M.; Thorpe, J. A. C. Photofrin PDT for early stage oesophageal cancer: long term results in 40 patients and literature review. Photodiag. Photodynam. Therap. 2009, 6, 159-166. [41] Douer, D.; Hu, W.; Giralt, S.; Lill, M.; DiPersio, J. Arsenic trioxide (Trisenox) therapy for acute promyelocytic leukemia in the setting of hematopoietic stem cell transplantation. Oncologist 2003, 8, 132-140. [42] Emmanouilides, C. Radioimmunotherapy for non-Hodgkin’s lymphoma. Semin. Oncol. 2003, 30, 531-544. [43] Anderson, P.; Nunez, R. Samarium lexidronam (Sm-153EDTMP): Skeletal radiation for osteoblastic bone metasteses and osteosarcoma. Expert Rev. Anticancer Therap. 2007, 7, 1517-1527. [44] Shirley, M.; McCormack, P. L. Radium-223 Dichloride: A review of its use in patients with castration-resistant prostate cancer with symptomatic bone metastases. Drugs 2014, 74, 579-586. [45] Windsor, P. M. Predictors of response to strontium-89 (Metastron1) in skeletal metastases from prostate cancer: report of a single centre’s 10-year experience. Clin. Oncol. 2001, 13, 219–222. [46] Rosenberg, B.; Vancamp, L.; Krigas, T. Inhibition of cell division in Escherichia coli by electrolysis products from a platinum electrode. Nature 1965, 205, 698-699. [47] Howell, S. B.; Safaei, R.; Larson, C. A.; Sailor, M. J. Copper transporters and the cellular pharmacology of the platinumcontaining cancer drugs. Mol. Pharmacol. 2010, 77, 887-894. [48] Fuertes, M. A.; Castilla, J.; Alonso, C.; Perez, J. M. Cisplatin biochemical mechanism of action: from cytotoxicity to induction of cell death through interconnections between apoptotic and necrotic pathways. Curr. Med. Chem. 2003, 10, 257-266. [49] Reedijk, J. Fast and slow versus strong and weak metal–DNA binding: consequences for anti-cancer activity. Metallomics 2012, 4, 628-632. [50] Casini, A.; Reedijk, J. Interactions of anticancer Pt compounds with proteins: an overlooked topic in medicinal inorganic chemistry? Chem. Sci. 2012, 3, 3135-3144. [51] Li, H.; Snelling, J. R.; Barrow, M. P.; Scrivens, J. H.; Sadler, P. J.; O’Connor, P. B. Mass spectrometric strategies to improve the identification of Pt(II)-modification sites on peptides and proteins. J. Am. Soc. Mass Spectrom. 2014, 25, 1217-1227. [52] Stordal, B.; Pavlakis, N.; Davey, R. Oxaliplatin for the treatment of cisplatin-resistant cancer: A systematic review. Cancer Treat. Rev. 2007, 33, 347-357. [53] Di Pasqua, A. J.; Goodisman, J.; Dabrowiak, J. C. Understanding how the platinum anticancer drug carboplatin works: From the bottle to the cell. Inorg. Chim. Acta 2012, 389, 29-35. [54] Umezawa, H.; Maeda, K.; Takeuchi, T.; Okami, Y. New antibiotics, bleomycin A and B. J. Antibiot., Ser. A 1966, 19, 200-209. [55] Burger, R. M.; Projan, S. J.; Horwitz, S. B.; Peisach, J. The DNA cleavage mechanism of iron-bleomycin. Kinetic resolution of strand scission from base propenal release. J. Biol. Chem. 1986, 261, 15955-15959. [56] Chin, J. Developing artificial hydrolytic metalloenzymes by a unified mechanistic approach. Acc. Chem. Res. 1991, 24, 145-152. [57] Cowan, J. A. Chemical nucleases. Curr. Opin. Chem. Biol. 2001, 5, 634-642. [58] Haner, R. Artificial ribonucleases. Chimia 2001, 55, 1035-1037. [59] Jin, Y.; Lewis, M. A.; Gokhale, N. H.; Long, E. C.; Cowan, J. A. Influence of stereochemistry and redox potentials on the single- and double-strand DNA cleavage efficiency of Cu(II)· and Ni(II)·Lys-Gly-His-derived ATCUN metallopeptides. J Am Chem Soc 2007, 129, 8353-8361. [60] Kikuta, E.; Aoki, S.; Kimura, E. A New Type of Potent Inhibitors of HIV-1 TAR RNA-Tat peptide binding by zinc(II)-macrocyclic tetraamine complexes. J. Am. Chem. Soc. 2001, 123, 7911-7912. [61] Livieri, M.; Mancin, F.; Tonellatoa, U.; Chin, J. Multiple functional group cooperation in phosphate diester cleavage promoted by Zn(II) complexes. Chem. Commun. 2004, 2862-63. [62] Morrow, J. R.; Amyes, T. L.; Richard, J. P. Phosphate binding energy and catalysis by small and large molecules. Acc. Chem. Res. 2008, 41, 539-548. [63] Sreedhara, A.; Cowan, J. A. Catalytic hydrolysis of DNA by metal ions and complexes. J. Biol. Inorg. Chem. 2001, 6, 337-347. [64] Suh, J. Synthetic artificial peptidases and nucleases using macromolecular catalytic systems. Acc. Chem. Res. 2003, 36, 562-570. [65] Suh, J.; Chei, W. S. Metal complexes as artificial proteases: toward catalytic drugs. Curr. Op. Chem. Biol. 2008, 12, 207-213. [66] Yashiro, M.; Ishikubo, A.; Komiyama, M. Dinuclear lanthanum(III) complex for Efficient hydrolysis of DNA. J. Biochem. 1996, 120, 1067-1069. [67] Yoo, S. H.; Lee, B. J.; Kim, H.; Suh, J. Artificial metalloprotease with active site comprising aldehyde group and Cu(II) cyclen complex. J. Am. Chem. Soc. 2005, 127, 9593-9602. [68] Hocharoen, L.; Cowan, J. A. Metallotherapeutics: Novel strategies in drug design. Chem. Eur. J. 2009, 15, 8670-8676. [69] Sarkar, B.; Wigfield, Y. Evidence for albumin-Cu(II)-amino acid ternary complex. Can. J. Biochem. 1968, 46, 601-607. [70] Lau, S.-J.; Kruck, T. P. A.; Sarkar, B. Peptide molecule mimicking the copper(II) transport site of human serum albumin. Comparative study between the synthetic site and albumin. J. Biol. Chem. 1974, 249, 5878-5884. [71] Brittain, I. J.; Huang, X.; Long, E. C. Selective recognition and cleavage of RNA loop structures by Ni (II)-Xaa-Gly-His metallopeptides. Biochemistry 1998, 37, 12113-12120. [72] Camerman, N.; Camerman, A.; Sarkar, B. Molecular design to mimic the copper(II) transport site of human albumin. The crystal and molecular structure of copper(II)-glycylglycyl-Lhistidine-N-methyl amide monoaquo complex. Can. J. Chem. 1976, 54, 1309-1316. [73] Conato, C.; Kozlowski, H.; Mlynarz, P.; Pulidori, F.; Remelli, M. Copper and nickel complex-formation equilibria with Lys-Gly-His-Lys, a fragment of the matricellular protein SPARC. Polyhedron 2002, 21, 1469-1474. [74] Harford, C.; Sarkar, B. Amino terminal Cu(II)- and Ni(II)-binding (ATCUN) motif of proteins and peptides: metal binding, DNA Unauthenticated Download Date | 6/15/17 8:02 PM cleavage, and other properties. Acc. Chem. Res. 1997, 30, 123-130. [75] McDonald, M. R.; Fredericks, F. C.; Margerum, D. W. Characterization of copper(III)-tetrapeptide complexes with histidine as the third residue. Inorg. Chem. 1997, 36, 3119-3124. [76] Gokhale, N. H.; Cowan, J. A. Metallopeptide-promoted inactivation of angiotensin-converting enzyme and endothelinconverting enzyme 1: toward dual-action therapeutics. J. Biol. Inorg. Chem. 2006, 11, 937-947. [77] Gokhale, N. H.; Bradford, S.; Cowan, J. A. Catalytic inactivation of human carbonic anhydrase I by a metallopeptidesulfonamide conjugate is mediated by oxidation of active site residues. J. Am. Chem. Soc. 2008, 130, 2388-2389. [78] Joyner, J. C.; Hocharoen, L.; Cowan, J. A. Targeted catalytic inactivation of angiotensin converting enzyme by lisinoprilcoupled transition metal chelates. J. Am. Chem. Soc. 2012, 134, 3396-3410. [79] Kim, M.-s.; Jeon, J. W.; Suh, J. Angiotensin-cleaving catalysts: conversion of N-terminal aspartate to pyruvate through oxidative decarboxylation catalyzed by Co(III)cyclen. J. Biol. Inorg. Chem. 2005, 10, 364-372. [80] Chei, W. S.; Ju, H.; Suh, J. New chelating ligands for Co(III)based peptide-cleaving catalysts selective for pathogenic proteins of amyloidoses. J. Biol. Inorg. Chem. 2011, 16, 511-519. [81] Podsiadlo, P.; Komiyama, T.; S., F. R.; Blum, O. Furin inhibition by compounds of copper and zinc. J. Biol. Chem. 2004, 279, 36219–36227. Past, Current and Future Use of Metals in Medicine 23 [82] Joyner, J. C.; Hodnick, W. F.; Cowan, A. S.; Tamuly, D.; Boyd, R.; Cowan, J. A. Antimicrobial metallopeptides with broad nuclease and ribonuclease activity. Chem. Commun. 2013, 49, 2118-2120. [83] Preihs, C.; Arambula, J. F.; Magda, D.; Jeong, H.; Yoo, D.; Cheon, J.; Siddik, Z. H.; Sessler, J. L. Recent developments in texaphyrin chemistry and drug discovery. Inorg. Chem. 2013, 52, 12184−12112. [84] Kameshima, W.; Ishizuka, T.; Minoshima, M.; Yamamoto, M.; Sugiyama, H.; Xu, Y.; Komiyama, M. Conjugation of peptide nucleic acid with a pyrrole/imidazole polyamide to specifically recognize and cleave DNA. Ang. Chem. Int. Ed. 2013, 52, 13681 –13684. [85] Joyner, J. C.; Keuper, K. D.; Cowan, J. A. Kinetics and mechanisms of oxidative cleavage of HIV RRE RNA by Rev-coupled transition metal–chelates. Chem. Sci. 2013. [86] Joyner, J. C.; Cowan, J. A. Targeted cleavage of HIV RRE RNA by Rev-coupled transition metal chelates. J. Am. Chem. Soc. 2011, 133, 9912-9922. [87] Bradford, S.; Cowan, J. A. Catalytic metalloDrugs targeting HCV IRES RNA. Chem. Commun. 2012, 48, 3118-3120. [88] Bradford, S. S.; Ross, M. J.; Fidai, I.; Cowan, J. A. Insight into the recognition, binding, and reactivity of catalytic metallodrugs targeting stem loop IIb of hepatitis C IRES RNA. ChemMedChem 2014, 9, 1275–1285. [89] Cowan, J. A. Catalytic metallodrugs. Pure & App. Chem. 2008, 80, 1799-1810. Unauthenticated Download Date | 6/15/17 8:02 PM