[Emerging Infectious Diseases * Volume 4 * Number 1 * January - March 1998] Synopses Proteases of Malaria Parasites: New Targets for Chemotherapy Philip J. Rosenthal San Francisco General Hospital and University of California, San Francisco, California, USA --------------------------------------------------------------------------- The increasing resistance of malaria parasites to antimalarial drugs is a major contributor to the reemergence of the disease as a major public health problem and its spread in new locations and populations. Among potential targets for new modes of chemotherapy are malarial proteases, which appear to mediate processes within the erythrocytic malarial life cycle, including the rupture and invasion of infected erythrocytes and the degradation of hemoglobin by trophozoites. Cysteine and aspartic protease inhibitors are now under study as potential antimalarials. Lead compounds have blocked in vitro parasite development at nanomolar concentrations and cured malaria-infected mice. This review discusses available antimalarial agents and summarizes experimental results that support development of protease inhibitors as antimalarial drugs. Hundreds of millions of cases of malaria occur annually, and infections with _Plasmodium falciparum_, the most virulent human malaria parasite, cause more than one million deaths per year (1). Despite extensive control efforts, the incidence of the disease is not decreasing in most malaria-endemic areas of the world, and in some it is clearly increasing (2). Malaria also remains a major risk to travelers from industrialized to developing countries. Because malaria parasites are increasingly resistant to antimalarial drugs, appropriately counseled travelers to malaria-endemic regions are more likely to contract malaria now than they were 40 years ago. Malaria control efforts include attempts to develop an effective vaccine, eradicate mosquito vectors, and develop new drugs (2,3). However, the development of a vaccine has proven very difficult, and a highly effective vaccine will probably not be available in the near future (4). Efforts to control _Anopheles_ mosquitoes have had limited success, although the use of insecticide-impregnated bed nets does appear to reduce malaria-related death rates (5). In addition, methods to replace natural vector populations with mosquitoes unable to support parasite development are under study and may contribute to malaria control in the long term (6). However, the current limitations of vaccine and vector control, as well as the increasing resistance of malaria parasites to existing drugs, highlight the continued need for new antimalarial agents. Established Antimalarial Drugs Antimalarial drugs have been used for centuries. Early natural products, including the bark of the cinchona tree in South America and extracts of the wormwood plant in China, were among the first effective antimicrobial agents to be used. Cinchona bark was used in Europe beginning in the 17th century, and upon its isolation from bark in 1820, quinine became widely used. In the last 50 years, extensive efforts, including the screening of hundreds of thousands of compounds, have led to the development of a number of effective synthetic antimalarial drugs. The most important of these, chloroquine, has been the mainstay of antimalarial chemotherapy for the last 50 years. The compound eradicates parasites rapidly, has minimal toxicity, is widely available at low cost throughout the world, and needs to be taken only once a week for chemoprophylaxis. However, resistance to chloroquine has been steadily increasing since the drug's initial use in South America and Southeast Asia in the late 1950s. Chloroquine resistance is now widespread in most _P. falciparum_ -endemic areas of the world (3). Thus, the use of chloroquine for presumptive treatment of falciparum malaria or for chemoprophylaxis is usually no longer appropriate (7). Moreover, resistance to chloroquine of _P. vivax_, the second most lethal human malaria parasite, is increasing in South Asia (8). No other antimalarial drug (9-12) is as efficacious and safe as chloroquine (Table 1). The best antimalarial drug for treating chloroquine-resistant falciparum malaria remains quinine (or intravenous quinidine), which is fairly toxic; quinine resistance is increasing in Southeast Asia, particularly in the border areas of Thailand (9). Amodiaquine, used to treat chloroquine-resistant malaria in developing countries, is also quite toxic, and resistance to it is also common (13). Mefloquine (14) is widely used for chemoprophylaxis against chloroquine-resistant _P. falciparum_, but its use is limited by toxicity (15) and (in the developing world) high cost. Mefloquine is not approved for treatment of malaria in the United States because of the neurotoxicity of doses required for the treatment. Fansidar, a combination of sulfadoxine and pyrimethamine, is no longer recommended for chemoprophylaxis because of its dermatologic toxicity (15). Fansidar is also not an ideal drug for treatment because it is slow acting, but it is increasingly important in treating chloroquine-resistant malaria in developing countries because economic constraints limit the use of other agents (16). The use of both mefloquine and Fansidar will increasingly be limited by drug resistance, already widespread in parts of Southeast Asia (9,17). Table 1. Established antimalarial drugs (sup a) --------------------------------------------------------------------------- Drug Role Best Feature(s) Limitations --------------------------------------------------------------------------- Chloroquine TX of and CP Very safe; low Widespread R against non- Pf cost; long and sensitive Pf half-life parasites Quinine/quinidine Best TX for Pf Limited R; Fairly toxic malaria; low cost rapidly acting (cinchonism, cardiac) Amodiaquine (sup TX of R Pf malaria Low cost Toxicity (bone b) marrow, liver); R common Mefloquine CP against R Pf Relatively little Moderately toxic malaria; not R, though (mostly CNS); approved for TX in increasing; long high cost; R in United States half-life SE Asia Fansidar TX of Pf malaria; Relatively low Skin toxicity (can no longer cost; long be fatal); recommended for CP half-life increasing R Primaquine Eradication of Only drug for Hemolysis with chronic liver this indication G6PD deficiency; stage Pv, Po increasing R malaria Proguanil (sup b) CP only (often Low cost; R common with chloroquine) nontoxic Maloprim (sup b) CP only (often Low cost R common; skin with chloroquine) rashes Tetracyclines CP; TX of Pf Low cost Skin and malaria in gastrointestinal combination with toxicity quinine --------------------------------------------------------------------------- (sup a) TX, therapy; CP, chemoprophylaxis; R, resistance/resistant; Pf, _Plasmodium falciparum_; Pv, _P. vivax_; Po, _P. ovale_; CNS, central nervous system; G6PD, glucose 6-phosphate dehydrogenase. (sup b) Not available in the United States. --------------------------------------------------------------------------- Other antimalarial drugs have specialized uses. Tetracyclines and some other antibiotics (clindamycin, sulfas) are slow acting and generally best used as an adjunct to quinine therapy in treating falciparum malaria (9). Doxycycline is also used for chemoprophylaxis in regions with high levels of drug resistance, especially Southeast Asia (10,17). Other drugs for chemoprophylaxis include proguanil, which remains effective in combination with chloroquine in many areas other than Southeast Asia, and Maloprim, a combination of dapsone and pyrimethamine (10,17). Resistance to these drugs is fairly common, however. Primaquine has a well-defined specific role: eradicating chronic liver stages of _P. vivax_ and _P. ovale_ after treating the acute blood infection with chloroquine. New Antimalarial Drugs Relatively few new antimalarial drugs are undergoing clinical testing (Table 2). Halofantrine, identified in the 1940s, was not developed until the 1980s; its use has been limited by variable oral absorption and cardiac toxicity (12,18). The drug is approved in the United States for treatment of chloroquine-resistant _P. falciparum_ infection, although in most cases quinine (or intravenous quinidine) is preferable. The most effective new drugs are artemisinin and related compounds. Artemisinin was isolated in 1972 from _Artemisia annua_, a plant used in China for centuries to treat fever (19). Artemisinin derivatives (artesunate, artelinate, artemether, arteether, dihydroartemisinin) have been synthesized and are undergoing extensive clinical testing. These compounds, which are already widely used in some areas, are potent, rapidly acting antimalarials that are effective against chloroquine-resistant _P. falciparum_ (20). Because recrudescences of infection after treatment are common, however, artemisinin and related compounds might best be used in combination with another drug. Table 2. New antimalarial drugs --------------------------------------------------------------------------- Drug Role Best Feature(s) Limitations -------------------------------------------------------------------------- Halofantrine TX of Pf malaria; Usually effective Variable not approved for against R Pf bioavailability, CP malaria cardiac toxicity Artemisinin and TX of Pf malaria Rapidly acting; Recurrence after related compounds effective against TX fairly common (sup a) multidrug-R strains Atovaquone ? TX of Pf Limited toxicity Limited studies so malaria; ? CP far show frequent (probably in recurrence after combination with TX proguanil) Pyronaridine (sup ? TX of Pf Effective against Studies limited to a) malaria R strains date Desferrioxamine ? TX of severe Pf Well tolerated Studies limited to malaria when used for date iron overload Azithromycin ? CP Limited toxicity Studies limited to date -------------------------------------------------------------------------- For abbreviations, see Table 1, footnote a. (sup a) Not available in the United States. Other compounds are under evaluation. Atovaqone (21), which is approved for treating patients with _Pneumocystis_ infections, appears to be effective against malaria in combination with proguanil (22), but its use has been limited by recrudescence after treatment. Pyronaridine, an acridine derivative used to treat malaria in China, has shown efficacy against falciparum malaria (23). The iron chelator desferrioxamine enhances the clearance of parasites in mild malaria (24) and, in conjunction with quinine and Fansidar, hastens recovery from deep coma in severe falciparum malaria (25). Azithromycin, a quinolone antibiotic, appears efficacious in malaria chemoprophylaxis (26). Malarial Proteases: New Targets for Chemotherapy The limitations of antimalarial chemotherapy underscore the need for new drugs, ideally directed against new targets. Potential targets for chemotherapy include malarial proteases (27). The erythrocytic life cycle, which is responsible for all clinical manifestations of malaria, begins when free merozoites invade erythrocytes. The intraerythrocytic parasites develop from small ring-stage organisms to larger, more metabolically active trophozoites and then to multinucleated schizonts. The erythrocytic cycle is completed when mature schizonts rupture erythrocytes, releasing numerous invasive merozoites. Proteases appear to be required for the rupture and subsequent reinvasion of erythrocytes by merozoite-stage parasites and for the degradation of hemoglobin by intraerythrocytic trophozoites figure. [fig] Figure. Protease targets in erythrocytic malaria parasites. The _Plasmodium falciparum_ erythrocytic life cycle is shown schematically, and data supporting cysteine (CP), serine (SP), and aspartic (AP) proteases of the different parasite stages as chemotherapeutic targets are provided in italics. Proteases and Erythrocyte Rupture and Invasion The rupture of erythrocytes by mature schizonts and the subsequent invasion of erythrocytes by free merozoites appear to require malarial protease activity, possibly to breach the erythrocyte cytoskeleton, a complex network of proteins. In addition, a number of malarial proteins are proteolytically processed during the late schizont and merozoite life-cycle stages; for example, merozoite surface protein-1 is processed in a manner inhibited by serine protease inhibitors (28), presumably to facilitate the complex series of events involved in erythrocyte rupture and invasion (29). Although the specific roles of different classes of proteases are not completely clear, inhibitors of cysteine and serine proteases have consistently blocked erythrocyte rupture and invasion (27). Candidate _P. falciparum_ rupture/invasion proteases have been identified, but none has been fully characterized biochemically or molecularly: 1) a 68 kD cysteine protease was identified in schizonts and merozoites and localized to the merozoite apex, suggesting that it may be released from the rhoptry organelle during invasion (30); 2) a cysteine protease of mature schizonts and a serine protease of merozoites were identified in highly synchronized parasites (31); 3) a serine protease was shown to be bound to the schizont/merozoite membrane by a glycosyl-phosphatidylinositol anchor, to be activated by phosphatidylinositol-specific phospholipase C during the merozoite stage, and to be capable of cleaving the erythrocyte cytoskeletal protein band 3 (32,33); 4) another protease, inhibited by both cysteine and serine protease inhibitors, hydrolyzed the erythrocyte cytoskeletal proteins spectrin and band 4.1 (34); and 5) the serine repeat antigen (35,36) and the related protein SERP H (37), both expressed in mature schizonts, have important similarities in their sequences with cysteine proteases. Further research should identify the specific biologic roles of the proteases mentioned and better characterize these enzymes, thus fostering the development of specific inhibitors. Host proteases may also play a role in erythrocyte rupture by _P. falciparum_. In recent studies, host urokinase was shown to bind to the surface of _P. falciparum_-infected erythrocytes, and the depletion of urokinase from parasite culture medium inhibited erythrocyte rupture by mature schizonts (38). This inhibition was reversed by exogenous urokinase. Drug Development Efforts Synthetic peptide inhibitors of the _P. falciparum_ schizont cysteine protease Pf 68 inhibited erythrocyte invasion by cultured parasites (39,40). The most effective peptide, GlcA-Val -Leu-Gly-Lys-NHC (sub 2)H(sub 5), inhibited the protease and blocked parasite development at high micromolar concentrations (40; Table 3). Although these results do not demonstrate levels of inhibition expected to be therapeutically relevant, they suggest that a specific protease activity is required for erythrocyte invasion by malaria parasites and thus is a potential target for antimalarial drugs. Proteases and Malarial Hemoglobin Degradation Extensive evidence suggests that the degradation of hemoglobin is necessary for the growth of erythrocytic malaria parasites, apparently to provide free amino acids for parasite protein synthesis (27,50). In _P. falciparum_, hemoglobin degradation occurs predominantly in trophozoites and early schizonts, the stages at which the parasites are most metabolically active. Trophozoites ingest erythrocyte cytoplasm and transport it to a large central food vacuole. In the food vacuole, hemoglobin is broken down into heme, a major component of malarial pigment (51), and globin, which is hydrolyzed to its constituent amino acids. The food vacuole is an acidic organelle analogous to lysosomes. Several lysosomal proteases are well characterized, including cysteine (cathepsins B, H, and L) and aspartic (cathepsin D) proteases (52), and malaria parasites contain analogous food vacuole proteases that degrade hemoglobin. At least two aspartic proteases and one cysteine protease have been isolated from purified _P. falciparum_ food vacuoles (53). Malarial aspartic protease activities have been identified (54-60). Two recently characterized aspartic proteases (plasmepsin I and plasmepsin II) are located in the food vacuole, have acid pH optima, and share sequence homology with other aspartic proteases (41,53,61,62). Furthermore, the aspartic proteases can cleave hemoglobin. One of the enzymes, plasmepsin I, cleaves native hemoglobin (53,59). Plasmepsin II appears to prefer denatured globin as a substrate (53). On the basis of these data, plasmepsin I is thought to be responsible for initial cleavages of hemoglobin after the molecule is transported to the food vacuole (53). Incubation of cultured _P. falciparum_ parasites with the protease inhibitor leupeptin caused trophozoite food vacuoles to fill with apparently undegraded erythrocyte cytoplasm (63-65). Analysis of the leupeptin-treated parasites showed that they contained large quantities of undegraded globin, while minimal globin was detectable in control parasites (64,66). Leupeptin inhibits both cysteine and some serine proteases, but the highly specific cysteine protease inhibitor E-64 also caused undegraded globin to accumulate. After parasites were incubated with inhibitors of other classes of proteases including the aspartic protease inhibitor pepstatin (63-67), globin did not accumulate. More recent studies that used nondenaturing electrophoretic methods demonstrated that cysteine protease inhibitors not only blocked malarial globin hydrolysis, but also inhibited earlier steps in hemoglobin degradation, including denaturation of the hemoglobin tetramer and the release of heme from globin (68). Another study showed that E-64, but not pepstatin, inhibited the production of hemozoin (the malarial end product of heme) by cultured parasites (69). These results suggest that a cysteine protease is required for initial steps in hemoglobin degradation by _P. falciparum_. A _P. falciparum trophozoite cysteine protease with biochemical features expected for a food vacuole hemoglobinase has been identified (31) and biochemically (70-72) and molecularly (73) characterized. This protease, called falcipain, degraded denatured and native hemoglobin in vitro; its acid pH optimum, substrate specificity, and inhibitor sensitivity indicated that it was a papain family cysteine protease (64,70,71). Specific inhibitors of falcipain blocked hemoglobin degradation and prevented parasite development. The degree of inhibition of falcipain by fluoromethyl ketones (44) and vinyl sulfones (46) correlated with their inhibition of hemoglobin degradation and parasite development, supporting the hypothesis that falcipain is the cysteine protease required for hemoglobin degradation. The specific mechanism for hemoglobin degradation in the malarial food vacuole remains unclear. As noted above, both the aspartic protease plasmepsin I and the cysteine protease falcipain have been identified in parasite food vacuoles and shown to cleave denatured and native hemoglobin in vitro (53,71). Results showing that only cysteine protease inhibitors block hemoglobin processing and globin hydrolysis in cultured parasites suggest that falcipain is required for initial steps of hemoglobin degradation (66-68,74). However, other studies have shown that native hemoglobin is cleaved by plasmepsin I, but not falcipain, in nonreducing conditions that may be present in the food vacuole (53,59,72). In any event, regardless of the exact sequence of hemoglobin processing, multiple enzymes, including at least the three proteases already identified, appear to participate in the degradation of hemoglobin. These proteases are thus logical targets for antimalarial drug development. Aminopeptidase activity has also been described in malaria parasites (75-77). This activity, with a neutral pH optimum, was not found in food vacuole lysates (77). When these lysates were incubated with hemoglobin, discrete peptide fragments, but not free amino acids, were identified (77). These results suggest that hemoglobin is degraded to small peptides in the food vacuole, that these peptides are transported to the parasite cytosol, and that additional processing of hemoglobin peptides is mediated by cytosolic aminopeptidase activity (77). Drug Development Efforts Both the cysteine protease inhibitor E-64 and the aspartic protease inhibitor pepstatin blocked _P. falciparum_ development (63-67). Administered together, the two inhibitors acted synergistically (67). However, only E-64 blocked globin hydrolysis (64-67). Numerous peptide-based cysteine protease inhibitors, including fluoromethyl ketones (44,70,78) and vinyl sulfones (46), inhibited falcipain at low nanomolar concentrations and inhibited _P. falciparum_ development and hemoglobin degradation at concentrations below 100 nanomolar (Table 3). In a malaria animal model, a fluoromethyl ketone that inhibited falcipain at low nanomolar concentrations blocked _P. vinckei_ protease activity in vivo after a single subcutaneous dose, and, when administered for 4 days, cured 80% of murine malaria infections (45). Thus, despite the theoretical limitations of potentially rapid degradation in vivo and inhibition of host proteases, peptide protease inhibitors show promise as candidate antimalarial drugs. Fluoromethyl ketones have subsequently shown toxicity in animal studies, but evaluations of related, apparently nontoxic inhibitors of falcipain as antimalarial drugs are under way. Table 3. Protease targets for chemotherapy --------------------------------------------------------------------------- Effective inhibitors (sup a) ------------------------------------------------------------------------ Protease Biologic Compound (Reference) In vitro (sup In vivo (sup role b) c) (mg/kg/day) (IC(sub 50); (Microgram M) ------------------------------------------------------------------------------------------------- Pf68 Erythrocyte GlcA-Val-Leu-Gly-Lys-NHC(sub 900 invasion 2)H(sub 5) (40) Plasmepsin Hemoglobin SC-50083 (41) 2-5 I degradation 0.25 Ro 40-4388 (42) Plasmepsin Hemoglobin Compound 7 (43) 20 II degradation Falcipain Hemoglobin Z-Phe-Arg-CH2F (44) 0.064 degradation Mu-Phe-HPh-CH2F (45) ~0.03 400 Mu-Leu-HPh-VSPh (46) 0.01 Oxalic bis ((2-hydroxy-1-naph- 7 thylmethylene)hydrazide) (47) 1-(2,5-dichlorophenyl)-3- 0.23 (4-quinolinyl)-2-propen-1-one (48) 7-chloro-1,2-dihydro-2-(2,3-di- 2 methoxy-phenyl)-5,5-dioxide- 4-(1H,10H)-phenothiazinone (49) ------------------------------------------------------------------------------------------------- (sup a) The structures of these compounds and details of the described studies are in the references noted. (sup b) Assays compared the development of new ring-form parasites or the uptake of [ (sup 3) H]hypoxanthine by treated and control parasites. (sup c) Cure of _Plasmodium vinckei_-infected mice. A computer model for the structure of falcipain was used to identify nonpeptide inhibitors (47). Screening of potential nonpeptide inhibitors identified a low micromolar lead compound (47; Table 3). Subsequent synthesis and testing of small molecules based on the structure of the lead compound have identified biologically active falcipain inhibitors, including chalcones that block parasite metabolism at submicromolar concentrations (48) and phenothiazines that block parasite metabolism and development at low micromolar concentrations (49). Peptidelike aspartic protease inhibitors are potent inhibitors of plasmepsins I and II. In independent studies SC-50083 (41), Ro 40-4388 (42), and "compound 7" (43) inhibited plasmepsin I or II at nanomolar concentrations and blocked parasite development at high nanomolar to micromolar concentrations (Table 3). Drug development efforts should be assisted by the recent determination of the structure of plasmepsin II (43). Inhibitors of aspartic and cysteine proteases have synergistic effects in inhibiting the growth of cultured malaria parasites (67), and these proteases also act synergistically to degrade hemoglobin in vitro (41). Therefore, the combination of inhibitors of malarial cysteine and aspartic proteases may provide the most effective chemotherapeutic regimen and best limit the development of parasite resistance to protease inhibitors. Ultimately, a better understanding of the biochemical properties and biologic roles of malarial proteases will foster the development of protease inhibitors that specifically inhibit parasite enzymes and thus are the most suitable candidates for chemotherapy. Acknowledgments Work in the author's laboratory was supported by grants from the National Institutes of Health, the UNDP/World Bank/WHO Special Programme for Research and Training in Tropical Diseases, and the American Heart Association. Dr. Rosenthal is associate professor, Department of Medicine, Division of Infectious Diseases, San Francisco General Hospital and the University of California, San Francisco. His research interests include the evaluation of proteases of malaria parasites as chemotherapeutic targets. Address for correspondence: Philip Rosenthal, Box 0811, University of California, San Francisco, CA 94143-0811, USA; fax: 415-206-6015; e-mail: rosnthl@itsa.ucsf.edu. References 1. Walsh JA. Disease problems in the Third World. Ann N Y Acad Sci 1989;569:1-16. 2. Oaks SC, Mitchell VS, Pearson GW, Carpenter CCJ, editors. Malaria: obstacles and opportunities. Washington: National Academy Press; 1991. 3. Olliaro P, Cattani J, Wirth D. Malaria, the submerged disease. JAMA 1996;275:230-3. 4. Hoffman SL, Miller LH. Perspectives on malaria vaccine development. In: Hoffman SL, editor. Malaria vaccine development. Washington: ASM Press; 1996. p. 1-13. 5. Alonso PL, Lindsay SW, Armstrong JRM, Conteh M, Hill AG, David PH, et al. The effect of insecticide-treated bed nets on mortality of Gambian children. Lancet 1991;337:1499-502. 6. Collins FH, Besansky NJ. Vector biology and the control of malaria in Africa. Science 1994;264:1874-5. 7. Centers for Disease Control. Recommendations for the prevention of malaria among travelers. JAMA 1990;263:2729-40. 8. Murphy GS, Basri H, Purnomo, Andersen EM, Bangs MJ, Mount DL, et al. Vivax malaria resistant to treatment and prophylaxis with chloroquine. Lancet 1993;341:96-100. 9. Panisko DM, Keystone JS. Treatment of malaria-1990. Drugs 1990;39:160-89. 10. Keystone JS. Prevention of malaria. Drugs 1990;39:337-54. 11. Wyler DJ. Malaria chemoprophylaxis for the traveler. N Engl J Med 1993;329:31-7. 12. White JJ. The treatment of malaria. N Engl J Med 1996;335:800-6. 13. Olliaro P, Nevill C, LeBras J, Ringwald P, Mussano P, Garner P, et al. Systematic review of amodiaquine treatment in uncomplicated malaria. Lancet 1996;348:1196-201. 14. Palmer KJ, Holliday SM, Brogden RN. Mefloquine: a review of its antimalarial activity, pharmacokinetic properties and therapeutic efficacy. Drugs 1993;45:430-75. 15. Luzzi GA, Peto TEA. Adverse effects of antimalarials: an update. Drug Saf 1993;8:295-311. 16. Hellgren U, Kihamia CM, Bergqvist Y, Lebbad M, Premji Z, Rombo L. Standard and reduced doses of sulfadoxine-pyrimethamine for treatment of _Plasmodium falciparum_ in Tanzania, with determination of drug concentrations and susceptibility _in vitro_. Trans R Soc Trop Med Hyg 1990;84:469-73. 17. Bradley DJ, Warhurst DC. Malaria prophylaxis: guidelines for travellers from Britain. BMJ 1995;310:709-14. 18. Bryson HM, Goa KL. Halofantrine: a review of its antimalarial activity, pharmacokinetic properties and therapeutic potential. Drugs 1992;43:236-58. 19. Meshnick SR, Taylor TE, Kamchonwongpaisan S. Artemisinin and the antimalarial endoperoxides: from herbal remedy to targeted chemotherapy. Microbiol Rev 1996;60:301-15. 20. de Vries PJ, Dien TK. Clinical pharmacology and therapeutic potential of artemisinin and its derivatives in the treatment of malaria. Drugs 1996;52:818-36. 21. Haile LG, Flaherty JF. Atovaquone: a review. Ann Pharmacother 1993;27:1488-94. 22. Radloff PD, Philipps J, Nkeyi M, Hutchinson D, Kremsner PG. Atovaquone and proguanil for _Plasmodium falciparum_ malaria. Lancet 1996;347:1511-4. 23. Ringwald P, Bickii J, Basco L. Randomised trial of pyronaridine versus chloroquine for acute uncomplicated falciparum malaria in Africa. Lancet 1996;347:24-8. 24. Gordeuk VR, Thuma PE, Brittenham GM, Biemba G, Zulu S, Simwanza G, et al. Iron chelation as a chemotherapeutic strategy for falciparum malaria. Am J Trop Med Hyg 1993;48:193-7. 25. Gordeuk V, Thuma P, Brittenham G, McLaren C, Parry D, Backenstose A, et al. Effect of iron chelation therapy on recovery from deep coma in children with cerebral malaria. N Engl J Med 1992;327:1473-7. 26. Anderson SL, Berman J, Kuschner R, Wesche D, Magill A, Wellde B, et al. Prophylaxis of _Plasmodium falciparum_ malaria with azithromycin administered to volunteers. Ann Intern Med 1995;123:771-3. 27. McKerrow JH, Sun E, Rosenthal PJ, Bouvier J. The proteases and pathogenicity of parasitic protozoa. Annu Rev Microbiol 1993;47:821-53. 28. Blackman MJ, Holder AA. Secondary processing of the _Plasmodium falciparum_ merozoite surface protein-1 (MSP1) by a calcium-dependent membrane-bound serine protease: shedding of MSP1(sub 33) as a noncovalently associated complex with other fragments of the MSP1. Mol Biochem Parasitol 1992;50:307-16. 29. Perkins ME. Erythrocyte invasion by the malarial merozoite: recent advances. Exp Parasitol 1989;69:94-9. 30. Bernard F, Schrev el J. Purification of a _Plasmodium berghei_ neutral endopeptidase and its localization in merozoite. Mol Biochem Parasitol 1987;26:167-74. 31. Rosenthal PJ, Kim K, McKerrow JH, Leech JH. Identification of three stage-specific proteinases of _Plasmodium falciparum_. J Exp Med 1987;166:816-21. 32. Braun-Breton C, Rosenberry TL, Pereira da Silva L. Induction of the proteolytic activity of a membrane protein in _Plasmodium falciparum_ by phosphatidyl inositol-specific phospholipase C. Nature 1988;332:457-9. 33. Roggwiller E, Bétoulle MEM, Blisnick T, Braun Breton C. A role for erythrocyte band 3 degradation by the parasite gp76 serine protease in the formation of the parasitophorous vacuole during invasion of erythrocytes by _Plasmodium falciparum_. Mol Biochem Parasitol 1996;82:13-24. 34. Deguercy A, Hommel M, Schrevel J. Purification and characterization of 37-kilodalton proteases from _Plasmodium falciparum_ and _Plasmodium berghei_ which cleave erythrocyte cytoskeletal components. Mol Biochem Parasitol 1990;38:233-44. 35. Knapp B, Hundt E, Nau U, Küpper HA. Molecular cloning, genomic structure and localization in a blood stage antigen of _Plasmodium falciparum_ characterized by a serine stretch. Mol Biochem Parasitol 1989;32:73-84. 36. Li W-B, Bzik DJ, Horii T, Inselburg J. Structure and expression of the _Plasmodium falciparum_ SERA gene. Mol Biochem Parasitol 1989;33:13-26. 37. Knapp B, Nau U, Hundt E, Küpper HA. A new blood stage antigen of _Plasmodium falciparum_ highly homologous to the serine-stretch protein SERP. Mol Biochem Parasitol 1991;44:1-14. 38. Roggwiller E, Fricaud A-C, Blisnick T, Braun-Breton C. Host urokinase-type plasminogen activator participates in the release of malaria merozoites from infected erythrocytes. Mol Biochem Parasitol 1997;86:49-59. 39. Schrevel J, Grellier P, Mayer R, Monsigny M. Neutral proteases involved in the reinvasion of erythrocytes by _Plasmodium_ merozoites. Biol Cell 1988;64:233-44. 40. Mayer R, Picard I, Lawton P, Grellier P, Barrault C, Monsigny M, et al. Peptide derivatives specific for a _Plasmodium falciparum_ proteinase inhibit the human erythrocyte invasion by merozoites. J Med Chem 1991;34:3029-35. 41. Francis SE, Gluzman IY, Oksman A, Knickerbocker A, Mueller R, Bryant ML, et al. Molecular characterization and inhibition of a _Plasmodium falciparum_ aspartic hemoglobinase. EMBO J 1994;13:306-17. 42. Moon RP, Tyas L, Certa U, Rupp K, Bur D, Jacquet C, et al. Expression and characterisation of plasmepsin I from _Plasmodium falciparum_. Eur J Biochem 1997;244:552-60. 43. Silva AM, Lee AY, Gulnik SV, Majer P, Collins J, Bhat TN, et al. Structure and inhibition of plasmepsin II, a hemoglobin-degrading enzyme from _Plasmodium falciparum_. Proc Natl Acad Sci U S A 1996;93:10034-9. 44. Rosenthal PJ, Wollish WS, Palmer JT, Rasnick D. Antimalarial effects of peptide inhibitors of a _Plasmodium falciparum_ cysteine proteinase. J Clin Invest 1991;88:1467-72. 45. Rosenthal PJ, Lee GK, Smith RE. Inhibition of a _Plasmodium vinckei_ cysteine proteinase cures murine malaria. J Clin Invest 1993;91:1052-6. 46. Rosenthal PJ, Olson JE, Lee GK, Palmer JT, Klaus JL, Rasnick D. Antimalarial effects of vinyl sulfone cysteine proteinase inhibitors. Antimicrob Agents Chemother 1996;40:1600-3. 47. Ring CS, Sun E, McKerrow JH, Lee GK, Rosenthal PJ, Kuntz ID, et al. Structure-based inhibitor design by using protein models for the development of antiparasitic agents. Proc Natl Acad Sci U S A 1993;90:3583-7. 48. Li R, Kenyon GL, Cohen FE, Chen X, Gong B, Dominguez JN, et al. _In vitro_ antimalarial activity of chalcones and their derivatives. J Med Chem 1995;38:5031-7. 49. Domínguez JN, López S, Charris J, Iarruso L, Lobo G, Semenov A, et al. Synthesis and antimalarial effects of phenothiazine inhibitors of a _Plasmodium falciparum_ cysteine protease. J Med Chem 1997;40:2726-32. 50. Rosenthal PJ, Meshnick SR. Hemoglobin catabolism and iron utilization by malaria parasites. Mol Biochem Parasitol 1996;83:131-9. 51. Slater AFG. Malaria pigment. Exp Parasitol 1992;74:362-5. 52. Bond JS, Butler PE. Intracellular proteases. Annu Rev Biochem 1987;56:333-64. 53. Gluzman IY, Francis SE, Oksman A, Smith CE, Duffin KL, Goldberg DE. Order and specificity of the _Plasmodium falciparum_ hemoglobin degradation pathway. J Clin Invest 1994;93:1602-8. 54. Gyang FN, Poole B, Trager W. Peptidases from _Plasmodium falciparum cultured in vitro. Mol Biochem Parasitol 1982;5:263-73. 55. Aissi E, Charet P, Bouquelet S, Biguet J. Endoprotease in _Plasmodium yoelii_ _nigeriensis_. Comp Biochem Physiol 1983;74B:559-66. 56. Sherman IW, Tanigoshi L. Purification of _Plasmodium lophurae_ cathepsin D and its effects on erythrocyte membrane proteins. Mol Biochem Parasitol 1983;8:207-26. 57. Sato K, Fukabori Y, Suzuki M. _Plasmodium berghei_: a study of globinolytic enzyme in erythrocytic parasite. Zbl Bakt Hyg A 1987;264:487-95. 58. Bailly E, Savel J, Mahouy G, Jaureguiberry G. _Plasmodium falciparum_: isolation and characterization of a 55-kDa protease with a cathepsin D-like activity from _P. falciparum_. Exp Parasitol 1991;72:278-84. 59. Goldberg DE, Slater AFG, Beavis R, Chait B, Cerami A, Henderson GB. Hemoglobin degradation in the human malaria pathogen _Plasmodium falciparum_: a catabolic pathway initiated by a specific aspartic protease. J Exp Med 1991;173:961-9. 60. Vander Jagt DL, Hunsaker LA, Campos NM, Scaletti JV. Localization and characterization of hemoglobin-degrading aspartic proteinases from the malarial parasite _Plasmodium falciparum_. Biochim Biophys Acta 1992;1122:256-64. 61. Dame JB, Reddy GR, Yowell CA, Dunn BM, Kay J, Berry C. Sequence, expression and modeled structure of an aspartic proteinase from the human malaria parasite _Plasmodium falciparum_. Mol Biochem Parasitol 1994;64:177-90. 62. Hill J, Tyas L, Phylip LH, Kay J, Dunn BM, Berry C. High level expression and characterisation of plasmepsin II, an aspartic proteinase from _Plasmodium falciparum_. FEBS Lett 1994;352:155-8. 63. Dluzewski AR, Rangachari K, Wilson RJM, Gratzer WB. _Plasmodium falciparum_: protease inhibitors and inhibition of erythrocyte invasion. Exp Parasitol 1986;62:416-22. 64. Rosenthal PJ, McKerrow JH, Aikawa M, Nagasawa H, Leech JH. A malarial cysteine proteinase is necessary for hemoglobin degradation by _Plasmodium falciparum_. J Clin Invest 1988;82:1560-6. 65. Vander Jagt DL, Caughey WS, Campos NM, Hunsaker LA, Zanner MA. Parasite proteases and antimalarial activities of protease inhibitors. Prog Clin Biol Res 1989;313:105-18. 66. Rosenthal PJ. _Plasmodium falciparum_: effects of proteinase inhibitors on globin hydrolysis by cultured malaria parasites. Exp Parasitol 1995;80:272-81. 67. Bailly E, Jambou R, Savel J, Jaureguiberry G. _Plasmodium falciparum_: differential sensitivity in vitro to E-64 (cysteine protease inhibitor) and pepstatin A (aspartyl protease inhibitor). Journal of Protozoology 1992;39:593-9. 68. Gamboa de Domínguez ND, Rosenthal PJ. Cysteine proteinase inhibitors block early steps in hemoglobin degradation by cultured malaria parasites. Blood 1996;87:4448-54. 69. Asawamahasakda W, Ittarat I, Chang C-C, McElroy P, Meshnick SR. Effects of antimalarials and protease inhibitors on plasmodial hemozoin production. Mol Biochem Parasitol 1994;67:183-91. 70. Rosenthal PJ, McKerrow JH, Rasnick D, Leech JH. _Plasmodium falciparum_: inhibitors of lysosomal cysteine proteinases inhibit a trophozoite proteinase and block parasite development. Mol Biochem Parasitol 1989;35:177-84. 71. Salas F, Fichmann J, Lee GK, Scott MD, Rosenthal PJ. Functional expression of falcipain, a _Plasmodium falciparum_ cysteine proteinase, supports its role as a malarial hemoglobinase. Infect Immun 1995;63:2120-5. 72. Francis SE, Gluzman IY, Oksman A, Banerjee D, Goldberg DE. Characterization of native falcipain, an enzyme involved in _Plasmodium falciparum_ hemoglobin degradation. Mol Biochem Parasitol 1996;83:189-200. 73. Rosenthal PJ, Nelson RG. Isolation and characterization of a cysteine proteinase gene of _Plasmodium falciparum_. Mol Biochem Parasitol 1992;51:143-52. 74. Kamchonwongpaisan S, Samoff E, Meshnick SR. Identification of hemoglobin degradation products in _Plasmodium falciparum_. Mol Biochem Parasitol 1997;86:179-86. 75. Vander Jagt DL, Baack BR, Hunsaker LA. Purification and characterization of an aminopeptidase from _Plasmodium falciparum_. Mol Biochem Parasitol 1984;10:45-54. 76. Curley GP, O'Donovan SM, McNally J, Mullally M, O'Hara H, Troy A, et al. Aminopeptidases from _Plasmodium falciparum_, _Plasmodium chabaudi_, and _Plasmodium berghei_. J Eukaryot Microbiol 1994;41:119-23. 77. Kolakovich KA, Gluzman IY, Duffin KL, Goldberg DE. Generation of hemoglobin peptides in the acidic digestive vacuole of _Plasmodium falciparum_ implicates peptide transport in amino acid production. Mol Biochem Parasitol 1997;87:123-35. 78. Rockett KA, Playfair JHL, Ashall F, Targett GAT, Angliker H, Shaw E. Inhibition of intraerythrocytic development of _Plasmodium falciparum_ by proteinase inhibitors. FEBS Lett 1990;259:257-9. --------------------------------------------------------------------------- Emerging Infectious Diseases National Center for Infectious Diseases Centers for Disease Control and Prevention Atlanta, GA URL: ftp://ftp.cdc.gov/pub/EID/vol4no1/ascii/rosenth.txt Please note that figures and equations are not available in ASCII format; their placement within the text is noted by [fig] and [eq], respectively. Greek symbols are spelled out. The following codes are used: (ft) for footnote; (sup) for superscript; (sub) for subscript; >/= for greater than or equal to. Italics are indicated by underlining before and after the italicized word(s) (e.g., _E.coli_ for E. coli).