TrypPROTACs Unlocking New Therapeutic Strategies for Chagas Disease
Abstract
:1. Chagas Disease: An Ongoing Public Health Challenge
2. Current Treatment and Limitations
3. Morphological Forms and Mitochondrial DNA Architecture in T. cruzi
4. Molecular Targets and Biological Pathways of T. cruzi
4.1. Lipid and Sterol Metabolism (Membrane Structure and Function)
4.1.1. Ergosterol Biosynthesis
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
20 | 40.17% squalene accumulation | 5.53432 | [82] | |
21 | 38.88% squalene accumulation | 5.18934 | [82] | |
22 | 37.94% squalene accumulation | 4.88092 | [82] | |
23 | IC50 = 0.8 µM (TcOSC) | 7.2436 | [84] | |
24 | IC50 = 0.47 µM (TcOSC) | 5.6504 | [85] | |
25 | IC50 = 0.55 µM (TcOSC) | 4.4767 | [86] | |
26 | IC50 = 0.80 µM (TcOSC) | 4.1773 | [86] | |
27 | IC50 = 42 nM (TcOSC) | 4.6691 | [87] | |
28 | IC50 = 44 nM (TcOSC) | 3.4200 | [87] |
4.1.2. Lipid Biosynthesis Pathway
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
29 | Ki < 25 nM (TcCYP51) | 6.4821 | [89] | |
30 | Kd < 2.2 µM (TcCYP51) | 8.7364 | [90] | |
31 | Kd = 8.4 nM (TcCYP51) | 6.8557 | [91] | |
32 | Kd < 10 nM (TcCYP51) | 5.2866 | [92] | |
33 | IC50 = 0.1 µM (TcCYP51) | 5.11372 | [93] | |
34 | >99% inhibition (at 3 µM) (Tc24-SMT) | 7.2633 | [94] | |
35 | IC50 = 116 nM (ScFTase) | 6.2462 | [96] | |
36 | IC50 = 125 nM (ScFTase) | 7.4451 | [96] | |
37 | — | 5.6755 | [98] |
4.2. Energy Metabolism and Nucleotide Synthesis
4.2.1. Glycolytic Pathway
Glyceraldehyde-3-Phosphate Dehydrogenase
Hexokinase/Glucokinase
Phosphofructokinase
4.2.2. Metabolism of Pentose Phosphate
6-Phosphogluconate Dehydrogenase
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
44 | Ki = 0.71 µM (TcGlcK) | −1.2873 | [113] | |
45 | Ki = 1.3 µM (TcGlcK) | −1.7591 | [113] | |
46 | Ki = 4.2 µM (TcGlcK) | −1.6591 | [113] | |
47 | Ki = 32 µM IC50 = 16.1 µM (TcGlcK) | −1.7837 | [113] | |
48 | Ki = 6.2 µM (TcGlcK) | 4.6077 | [114] | |
49 | Ki = 4.8 µM (TcGlcK) | 5.5421 | [114] | |
50 | Ki = 5.5 µM (TcGlcK) | 4.2176 | [115] | |
51 | IC50 = 0.041 µM (TcPFK) | 2.498 | [118] | |
52 | IC50 = 0.052 µM (TcPFK) | 2.2904 | [118] | |
53 | Ki = 0.16 µM (T. brucei 6PGDH) | −2.0979 | [121] |
4.2.3. Synthesis of Nucleotides
Purine Phosphoribosyltransferases
Dihydrofolate Reductase
Pteridine Reductase
Dihydroorotate Dehydrogenase
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
54 | Ki = 0.6 µM (TcHGXPRT) Ki = 0.006 µM (TcHGPRT) | −1.1483 | [124] | |
55 | Ki = 1.6 µM (TcHGXPRT) Ki = 0.004 µM (TcHGPRT) | −0.7305 | [124] | |
56 | Ki = 0.7 µM (TcHGXPRT) Ki = 1.4 µM (TcHGPRT) | −0.4349 | [124] | |
57 | Ki = 1.3 nM (TcDHFR–TS) | 3.0578 | [128] | |
58 | Ki = 7 µM (TcPTR1) | 1.2215 | [132] | |
59 | Ki = 0.11 µM (TcPTR1) | 0.2684 | [132] | |
60 | Kiapp = 0.024 µM (TcDHODH) | −4.8302 | [135] | |
61 | Kiapp = 0.033 µM IC50~33 nM (TcDHODH) | −4.8302 | [135,136] | |
62 | IC50~46 nM (TcDHODH) | 1.7084 | [136] |
4.3. Oxidative Stress Defense Mechanisms
Metabolism Dependent on Thiol Groups
Trypanothione Reductase
Trypanothione Synthetase
Tryparedoxin Peroxidase
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
63 | Ki = 330 nM (TcTR) | 6.5832 | [141] | |
64 | Ki = 510 nM (TcTR) | 10.571 | [142,143] | |
65 | Ki = 2.0 µM (TcTR) | 3.3366 | [144] | |
66 | DS = −6.247 kcal/mol (TcTR) | 6.2142 | [145] | |
67 | Inhibition (at 30 µM) = 65.4% (TcTryS) | 2.309 | [148] | |
68 | IC50 = 2.21 µM (T. cruzi) | 1.533 | [151] | |
69 | IC50 = 2.74 µM (T. cruzi) | 2.3674 | [151] |
4.4. Protein Kinases and Signal Transduction
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
70 | EC50 = 1.85 µM (T. cruzi—amastigote) Binding Score = −8.3 kcal/mol (TcAkt) | 7.4842 | [153,154] | |
71 | IC50 = 1.55 μM (T. cruzi—epimastigote) IC50 = 421 nM (T. cruzi—trypomastigote) | 3.5408 | [155] | |
72 | IC50 = 2.42 μM (T. cruzi—epimastigote) IC50 = 578 nM (T. cruzi—trypomastigote) | 3.9692 | [155] | |
73 | — | 2.4084 | [156] | |
74 | IC50 = 0.017 μM (T. cruzi—epimastigote) | 1.9132 | [157] | |
75 | IC50 = 0.24 μM (T. cruzi—epimastigote) | 4.323 | [157] |
4.5. DNA Replication and Epigenetic Regulation
4.5.1. DNA Topoisomerases
4.5.2. Bromodomain Proteins
Bromodomain Factor 3
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
76 | GI50 = 38.5 µM (Sylvio X10 strain) GI50 = 24.6 µM (Esmeraldo strain) | 1.0203 | [159] | |
77 | EC50 = 78 nM (Topo IIα) | 2.91208 | [160] | |
78 | EC50 = 69 nM (Topo IIα) | 2.27697 | [160] | |
79 | EC50 = 87 nM (Topo IIα) | 2.77298 | [160] | |
80 | IC50 = 2.4 µM Kd = 1.7 µM (TcBDF3) | 2.2262 | [164] | |
81 | IC50 = 8.4 µM Kd = 4.0 µM (TcBDF3) | 3.1793 | [164] | |
82 | IC50 = 10.5 µM Kd = 4.8 µM (TcBDF3) | 2.5504 | [164] |
4.6. Host-Parasite Interaction and Immune Evasion
4.6.1. Transfer of Sialic Acid
Trans-Sialidase
4.6.2. Proteases (Cellular Invasion and Immune Escape)
Cysteine Protease (Cruzain)
4.7. Drug Targets and Resistance Mechanisms
Nitroreductases
Nitroreductase Type I
5. PROTACs: Mechanism of Action, Therapeutic Applications, and Key Challenges
5.1. Mechanism of Action
5.2. Therapeutic Applications of PROTACs
5.2.1. PROTACs in Oncology: Targeting Oncogenic Proteins
5.2.2. PROTACs for Antibacterial Drug Discovery
5.2.3. PROTACs in Antiviral Therapies
5.3. Advantages and Challenges of PROTACs in Drug Development
6. Molecular Design of TrypPROTACs for T. cruzi
6.1. Prioritized Targets and Representative Ligands
Target | Ligand | Affinity Data | LogP | Target Rationale | Comment |
---|---|---|---|---|---|
TcCYP51 | (31) | Kd = 8.4 nM | 6.85 | Essential for ergosterol biosynthesis and membrane integrity. Active across life stages. Resistance to azoles reported. Intracellular and structurally accessible. | Potent but highly lipophilic ligands; elevated LogP may impair solubility and passive diffusion. Requires optimized linker and solubility-enhancing strategies. |
(32) | Kd < 10.0 nM | 5.28 | |||
TcTR | (63) | Ki = 330 nM | 6.58 | Key redox enzyme with no human homolog. Essential under oxidative stress. Cytosolic and suitable for UPS access. | Compact soluble target; high ligand lipophilicity and globular shape may limit ternary complex formation. Demands polarity-balanced linker design. |
(64) | Ki = 510 nM | 10.57 | |||
TcAkt | (71) | IC50 = 421 nM | 3.54 | Involved in signaling, proliferation, and differentiation. Structurally divergent. Cytoplasmic and PROTAC-compatible despite no current precedents. | Favorable LogP and cellular accessibility; minimal structural modification required for PROTAC conversion. |
Hsp90 | (74) | IC50 = 0.017 µM | 1.91 | Essential chaperone for stress response and differentiation. Cytosolic, ligand-accessible, and successfully degraded by PROTACs in human cells. | Known PROTAC warhead; high-affinity ligand with optimal LogP and established use in mammalian systems. |
TcTopoIIα | (77) | IC50 = 78 nM | 2.91 | Crucial for DNA replication and repair. Nuclear, with high turnover and covalent ligands supporting sustained degradation. | Potent ligands with ideal LogP. Covalent binding (Cys477) may enhance degradation efficiency. |
(78) | IC50 = 78 nM | 2.27 | |||
TcBDF3 | (80) | IC50 = 2.4 µM Kd = 1.7 µM | 2.23 | Bromodomain protein regulating differentiation. Nuclear, structurally tractable, and analogous to known PROTAC targets like BRD4. | LogP compatible with permeability; hydrophobic pocket favors linker attachment. |
(81) | IC50 = 8.4 µM Kd = 4.0 µM | 3.27 | |||
Cruzain | (91) | IC50 = 0.081 nM | 5.04 | Major protease involved in immune evasion and differentiation. Immature form is cytosolic. Resistance reported; ligands available. | Excellent potency, but high LogP may limit solubility and selectivity. Needs careful linker and E3 ligase optimization. |
(95) | IC50 = 0.01 µM | 5.19 |
6.2. Comparative Exploration of the Ubiquitination Machinery in T. cruzi and Humans
6.3. Rational Linker Design in PROTACs
6.4. Experimental Validation of TrypPROTACs
7. Future Perspectives and Conclusions
Author Contributions
Funding
Institutional Review Board Statement
Informed Consent Statement
Data Availability Statement
Acknowledgments
Conflicts of Interest
Abbreviations
6PGDH | 6-Phosphogluconate dehydrogenase |
24-SMT | Δ24(25)-sterol methyltransferase |
AEK1 | AGC essential kinase 1 |
Akt | akt-like kinase |
BDF3 | bromodomain Factor 3 |
BRD4 | bromodomain-containing protein 4 |
BZN | benznidazole |
CC50 | half-maximal cytotoxic concentration |
CDK4/6 | cyclin-dependent kinases 4 and 6 |
CETSA | Cellular Thermal Shift Assay |
ClpC-ClpP complex | caseinolytic protease system |
CoMFA | Comparative Molecular Field Analysis |
CPx | cytosolic tryparedoxin peroxidase |
CRBN | Cereblon |
CRL | Cullin-RING Ligase |
CYP51 | sterol 14α-demethylase |
DHFR | dihydrofolate reductase |
DHFR–TS | dihydrofolate reductase–thymidylate synthase (bifunctional) |
DHODH | dihydroorotate dehydrogenase |
DS | docking score |
DUB | deubiquitinase |
EC50 | half-maximal effective concentration |
EGFR | epidermal growth factor receptor |
FMN | flavin mononucleotide |
FPP | farnesyl diphosphate synthase |
FPPS | farnesylpyrophosphate synthase |
FTase | farnesyltransferase |
G3P | glyceraldehyde-3-phosphate |
G6P | glucose-6-phosphate |
GAPDH | glyceraldehyde-3-phosphate dehydrogenase |
GI50 | half-maximal growth inhibitory concentration |
GlcK | glucokinase |
GR | glutathione reductase |
gRNAs | guide RNAs |
HER2 | human epidermal growth factor receptor 2 |
HGPRT | hypoxanthine-guanine phosphoribosyltransferase |
HGXPRT | hypoxanthine-guanine-xanthine phosphoribosyltransferase |
HiBiT | High-affinity Bioluminescent Tag |
HIV-1 | Human Immunodeficiency Virus type 1 |
HMG-CoA | 3-hydroxy-3-methylglutaryl coenzyme A |
HPV | Human Papillomavirus |
HQSAR | hologram quantitative structure–activity relationship |
Hsps | heat shock proteins |
HxK | hexokinase |
IAP | inhibitor of apoptosis protein |
IC50 | half-maximal inhibitory concentration |
IPC | inositolphosphorylceramide |
kDNA | kinetoplast DNA |
Kd | dissociation constant |
Ki | inhibition constant |
Kiapp | apparent inhibition constant |
Km | Michaelis–Menten constant |
Keap1 | Kelch-like ECH-associated protein 1 |
Kobs | observed rate constant |
KKT3 | kinetochore kinesin 3 |
L. mexicana | Leishmania mexicana |
LC50 | half-maximal lethal concentration |
LC–MS/MS | liquid chromatography–tandem mass spectrometry |
LogP | logarithm of the partition coefficient (octanol/water) |
LS | lanosterol synthase |
MDM2 | mouse double minute 2 homolog |
MPx | mitochondrial tryparedoxin peroxidase |
MuNANA | 4-methylumbelliferyl-N-acetylneuraminic acid assay |
MYC | myelocytomatosis oncogene |
NADPH | nicotinamide adenine dinucleotide phosphate (reduced form) |
NanoBiT | nanoluciferase binary technology |
NanoBRET | bioluminescence resonance energy transfer–based assays |
NanoLuc | nanoluciferase |
NFX | nifurtimox |
NMR | Nuclear Magnetic Resonance |
NRF2 | nuclear factor erythroid 2–related factor 2 |
NTR | nitroreductase |
OSC | oxidosqualene cyclase |
PFK | phosphofructokinase |
PH | pleckstrin homology |
POI | protein of interest |
PROTAC | proteolysis-targeting chimeras |
PRPP | 5-phospho-α-D-ribose 1-pyrophosphate |
PRT | purine phosphoribosyltransferase |
PTR | pteridine reductase |
QFIT | quantitative pharmacophore fit value |
qPCR | quantitative polymerase chain reaction |
QSAR | quantitative structure–activity relationship |
Rbx1 | RING-box protein |
ROS | reactive oxygen species |
RTK | receptor tyrosine kinase |
SAPA | shed acute-phase antigen |
SAR | structure–activity relationship |
SCF complex | Skp1–Cullin1–F-box |
SE | squalene epoxidase |
SEC31 | secretory protein 31 |
SI | selectivity index |
SOD | superoxide dismutase |
SQS | squalene synthase |
SPRING | small PROTAC-influenced RING-type |
T. brucei | Trypanosoma brucei |
T. cruzi | Trypanosoma cruzi |
TMQ | trimetrexate |
Topo3α | topoisomerase 3α |
TPD | targeted protein degradation |
TPSA | topological polar surface area |
TR | trypanothione reductase |
TS | trans-sialidase |
TryS | trypanothione synthetase |
TXNI | tryparedoxin I |
TXNPx | tryparedoxin peroxidase |
UPS | ubiquitin-proteasome system |
VHL | Von Hippel-Lindau |
References
- World Health Organization. Chagas Disease (Also Known as American Trypanosomiasis). 2025. Available online: https://www.who.int/news-room/fact-sheets/detail/chagas-disease-(american-trypanosomiasis) (accessed on 9 April 2025).
- Chagas, C. Nova tripanozomiaze humana: Estudos sobre a morfolojia e o ciclo evolutivo do Schizotrypanum cruzi n. gen., n. sp., ajente etiolojico de nova entidade morbida do homem. Mem. Inst. Oswaldo Cruz 1909, 1, 159–218. [Google Scholar] [CrossRef]
- Cucunubá, Z.M.; Gutiérrez-Romero, S.A.; Ramírez, J.-D.; Velásquez-Ortiz, N.; Ceccarelli, S.; Parra-Henao, G.; Henao-Martínez, A.F.; Rabinovich, J.; Basáñez, M.-G.; Nouvellet, P.; et al. The epidemiology of Chagas disease in the Americas. Lancet Reg. Health Am. 2024, 37, 100881. [Google Scholar] [CrossRef] [PubMed]
- Martins-Melo, F.R.; Ramos, A.N.; Alencar, C.H.; Heukelbach, J. Prevalence of Chagas disease in Brazil: A systematic review and meta-analysis. Acta Trop. 2014, 130, 167–174. [Google Scholar] [CrossRef]
- Cruz, D.S.; Damasceno, R.F.; Leite, S.F.; Cardoso, M.D.; Almeida, D.N.M.; de Souza, A.B.; de Jesus Santos, A.C.; Veira, T.M.; Ribeiro, A.L.P.; de Oliveira, L.C.; et al. Prevalence analysis of Chagas disease by age group in an endemic region of Brazil: Possible scenario of active vectorial transmission. IJID Reg. 2024, 12, 100400. [Google Scholar] [CrossRef]
- Navarro, M.; Reguero, L.; Subirà, C.; Blázquez-Pérez, A.; Requena-Méndez, A. Estimating Chagas disease prevalence and number of underdiagnosed and undertreated individuals in Spain. Travel Med. Infect. Dis. 2022, 47, 102284. [Google Scholar] [CrossRef]
- Carbajal-De-la-Fuente, A.L.; Sánchez-Casaccia, P.; Piccinali, R.V.; Provecho, Y.; Salvá, L.; Meli, S.; Cano, F.; Hernández, R.; Nattero, J. Urban vectors of Chagas disease in the American continent: A systematic review of epidemiological surveys. PLoS Neglected Trop. Dis. 2022, 16, e0011003. [Google Scholar] [CrossRef]
- Santos, F.; Magalhães-Junior, J.T.; Carneiro, I.O.; Santos, F.L.N.; Silva, Â.A.O.; Novais, J.M.C.B.; Santos, J.S.S.; Ribeiro, G., Jr.; Reis, M.G.; Franke, C.R. Eco-epidemiology of vectorial Trypanosoma cruzi transmission in a region of northeast Brazil. Acta Trop. 2022, 225, 106184. [Google Scholar] [CrossRef] [PubMed]
- Coura, J.R. The main sceneries of Chagas disease transmission. The vectors, blood and oral transmissions—A comprehensive review. Mem. Inst. Oswaldo Cruz 2015, 110, 277. [Google Scholar] [CrossRef]
- Boukaabar, M.; Oduro, B.; Chataa, P. Congenital transmission Chagas disease: The role of newborn therapy on the disease’s dynamics. PLoS ONE 2024, 19, e0308136. [Google Scholar] [CrossRef]
- de Sousa, D.R.T.; de Oliveira Guerra, J.A.; Ortiz, J.V.; do Nascimento Couceiro, K.; da Silva e Silva, M.R.H.; Jorge Brandão, A.R.; Guevara, E.; Arcanjo, A.R.L.; de Oliveira Júnior, E.F.; Smith-Doria, S.; et al. Acute Chagas disease associated with ingestion of contaminated food in Brazilian western Amazon. Trop. Med. Int. Health 2023, 28, 541–550. [Google Scholar] [CrossRef]
- Angheben, A.; Boix, L.; Buonfrate, D.; Gobbi, F.; Bisoffi, Z.; Pupella, S.; Gandini, G.; Aprili, G. Chagas disease and transfusion medicine: A perspective from non-endemic countries. Blood Transfus. 2015, 13, 540. [Google Scholar] [CrossRef]
- Contreras, G.A.; Golovko, G. Prevalence of Trypanosoma cruzi infection in solid organ transplant recipients: A neglected disease in America. Open Forum Infect. Dis. 2024, 11, ofae650. [Google Scholar] [CrossRef] [PubMed]
- Pérez-Molina, J.A.; Molina, I. Chagas disease. Lancet 2018, 391, 82–94. [Google Scholar] [CrossRef] [PubMed]
- Hochberg, N.S.; Montgomery, S.P. In the Clinic®: Chagas Disease. Ann. Intern. Med. 2023, 176, ITC17. [Google Scholar] [CrossRef]
- Rassi, A.; Rassi, A.; Marcondes de Rezende, J. American Trypanosomiasis (Chagas Disease). Infect. Dis. Clin. N. Am. 2012, 26, 275–291. [Google Scholar] [CrossRef]
- Santos, É.; Menezes Falcão, L. Chagas cardiomyopathy and heart failure: From epidemiology to treatment. Rev. Port. Cardiol. 2020, 39, 279–289. [Google Scholar] [CrossRef]
- Saraiva, R.M.; Mediano, M.F.F.; Mendes, F.S.N.S.; Sperandio da Silva, G.M.; Veloso, H.H.; Sangenis, L.H.C.; da Silva, P.S.; Mazzoli-Rocha, F.; Sousa, A.S.; Holanda, M.T.; et al. Chagas heart disease: An overview of diagnosis, manifestations, treatment, and care. World J. Cardiol. 2021, 13, 654. [Google Scholar] [CrossRef]
- Simões, M.V.; Romano, M.M.D.; Schmidt, A.; Martins, K.S.M.; Marin-Neto, J.A. Chagas Disease Cardiomyopathy. Int. J. Cardiovasc. Sci. 2018, 31, 173–189. [Google Scholar] [CrossRef]
- Andrade, M.V.; Noronha, K.V.M.D.S.; de Souza, A.; Motta-Santos, A.S.; Braga, P.E.F.; Bracarense, H.; de Miranda, M.C.C.; Nascimento, B.R.; Molina, I.; Martins-Melo, F.R.; et al. The economic burden of Chagas disease: A systematic review. PLoS Negl. Trop. Dis. 2023, 17, e0011757. [Google Scholar] [CrossRef] [PubMed]
- Lee, B.Y.; Bacon, K.M.; Bottazzi, M.E.; Hotez, P.J. Global economic burden of Chagas disease: A computational simulation model. Lancet Infect. Dis. 2013, 13, 342–348. [Google Scholar] [CrossRef]
- Gómez-Ochoa, S.A.; Rojas, L.Z.; Echeverría, L.E.; Muka, T.; Franco, O.H. Global, Regional, and National Trends of Chagas Disease from 1990 to 2019: Comprehensive Analysis of the Global Burden of Disease Study. Glob. Heart 2022, 17, 1150. [Google Scholar] [CrossRef] [PubMed]
- Kalil-Filho, R. Globalization of Chagas Disease Burden and New Treatment Perspectives. J. Am. Coll. Cardiol. 2015, 66, 1190–1192. [Google Scholar] [CrossRef] [PubMed]
- Pinto Dias, J.C. Human Chagas Disease and Migration in the Context of Globalization: Some Particular Aspects. J. Trop. Med. 2013, 2013, 789758. [Google Scholar] [CrossRef] [PubMed]
- Forsyth, C.J.; Hernandez, S.; Olmedo, W.; Abuhamidah, A.; Traina, M.I.; Sanchez, D.R.; Soverow, J.; Meymandi, S.K. Safety Profile of Nifurtimox for Treatment of Chagas Disease in the United States. Clin. Infect. Dis. 2016, 63, 1056. [Google Scholar] [CrossRef]
- Steverding, D. The history of Chagas disease. Parasites Vectors 2014, 7, 317. [Google Scholar] [CrossRef]
- Crespillo-Andújar, C.; Comeche, B.; Hamer, D.H.; Arevalo-Rodriguez, I.; Alvarez-Díaz, N.; Zamora, J.; Pérez-Molina, J.A. Use of benznidazole to treat chronic Chagas disease: An updated systematic review with a meta-analysis. PLoS Negl. Trop. Dis. 2022, 16, e0010386. [Google Scholar] [CrossRef]
- Altcheh, J. Randomized trial of benznidazole for chronic Chagas’ cardiomyopathy. Arch. Argent. Pediatr. 2016, 114, e124–e125. [Google Scholar] [CrossRef]
- Jackson, Y.; Wyssa, B.; Chappuis, F. Tolerance to nifurtimox and benznidazole in adult patients with chronic Chagas’ disease. J. Antimicrob. Chemother. 2020, 75, 690–696. [Google Scholar] [CrossRef]
- Crespillo-Andújar, C.; Chamorro-Tojeiro, S.; Norman, F.; Monge-Maillo, B.; López-Vélez, R.; Pérez-Molina, J.A. Toxicity of nifurtimox as second-line treatment after benznidazole intolerance in patients with chronic Chagas disease: When available options fail. Clin. Microbiol. Infect. 2018, 24, 1344.e1–1344.e4. [Google Scholar] [CrossRef]
- Francisco, A.F.; Jayawardhana, S.; Olmo, F.; Lewis, M.D.; Wilkinson, S.R.; Taylor, M.C.; Kelly, J.M. Challenges in Chagas Disease Drug Development. Molecules 2020, 25, 2799. [Google Scholar] [CrossRef]
- Aldasoro, E.; Posada, E.; Requena-Méndez, A.; Calvo-Cano, A.; Serret, N.; Casellas, A.; Sanz, S.; Soy, D.; Pinazo, J.; Gascon, J. What to Expect and When: Benznidazole Toxicity in Chronic Chagas’ Disease Treatment. J. Antimicrob. Chemother. 2018, 73, 1060–1067. [Google Scholar] [CrossRef]
- Bermudez, J.; Davies, C.; Simonazzi, A.; Real, J.P.; Palma, S. Current Drug Therapy and Pharmaceutical Challenges for Chagas Disease. Acta Trop. 2016, 156, 1–16. [Google Scholar] [CrossRef] [PubMed]
- Molina, I.; Salvador, F.; Sánchez-Montalvá, A.; Treviño, B.; Serre, N.; Avilés, A.S.; Almirante, B. Toxic Profile of Benznidazole in Patients with Chronic Chagas Disease: Risk Factors and Comparison of the Product from Two Different Manufacturers. Antimicrob. Agents Chemother. 2015, 59, 6125–6131. [Google Scholar] [CrossRef] [PubMed]
- Grossmann, U.; Rodriguez, M.-L. Chagas Disease Treatment Efficacy Markers: Experiences from a Phase III Study with Nifurtimox in Children. Front. Parasitol. 2023, 2, 1229467. [Google Scholar] [CrossRef]
- Mazzeti, A.L.; Capelari-Oliveira, P.; Bahia, M.T.; Mosqueira, V.C.F. Review on Experimental Treatment Strategies Against Trypanosoma cruzi. J. Exp. Pharmacol. 2021, 13, 409–432. [Google Scholar] [CrossRef] [PubMed]
- Porta, E.O.J.; Kalesh, K.; Steel, P.G. Navigating Drug Repurposing for Chagas Disease: Advances, Challenges, and Opportunities. Front. Pharmacol. 2023, 14, 1233253. [Google Scholar] [CrossRef]
- Morillo, C.A.; Waskin, H.; Sosa-Estani, S.; Bangher, M.C.; Cuneo, C.; Milesi, R.; Mallagray, M.; Apt, W.; Beloscar, J.; Gascon, J.; et al. Benznidazole and Posaconazole in Eliminating Parasites in Asymptomatic T. cruzi Carriers: The STOP-CHAGAS Trial. J. Am. Coll. Cardiol. 2017, 69, 939–947. [Google Scholar] [CrossRef]
- Roig-Sanchis, J.; Bosch-Nicolau, P.; Silgado, A.; Salvador, F.; Sánchez-Montalvá, A.; Aznar, M.; Oliveira, I.; Espinosa-Pereiro, J.; Serre-Delcor, N.; Pou, D.; et al. Long-Term Follow-Up of Individuals with Chagas Disease Treated with Posaconazole and Benznidazole in a Non-Endemic Region: The CHAGASAZOL Cohort. Clin. Microbiol. Infect. 2025, in press. [Google Scholar] [CrossRef]
- Torrico, F.; Gascón, J.; Ortiz, L.; Pinto, J.; Rojas, G.; Palacios, A.; Barreira, F.; Blum, B.; Schijman, A.G.; Vaillant, M.; et al. A Phase 2, Randomized, Multicenter, Placebo-Controlled, Proof-of-Concept Trial of Oral Fexinidazole in Adults with Chronic Indeterminate Chagas Disease. Clin. Infect. Dis. 2022, 76, e1186. [Google Scholar] [CrossRef]
- Pinazo, M.J.; Forsyth, C.; Losada, I.; Esteban, E.T.; García-Rodríguez, M.; Villegas, M.L.; Molina, I.; Crespillo-Andújar, C.; Gállego, M.; Ballart, C.; et al. Efficacy and safety of fexinidazole for treatment of chronic indeterminate Chagas disease (FEXI-12): A multicentre, randomised, double-blind, phase 2 trial. Lancet Infect. Dis. 2024, 24, 395–403. [Google Scholar] [CrossRef]
- Sangenito, L.S.; Branquinha, M.H.; Santos, A.L.S. Funding for Chagas disease: A 10-year (2009–2018) survey. Trop. Med. Infect. Dis. 2020, 5, 88. [Google Scholar] [CrossRef] [PubMed]
- Drugs for Neglected Diseases Initiative (DNDi). Study Shows Dramatically Shorter Treatment for Chagas Disease Could Be Just as Effective, and Significantly Safer. DNDi Organization. 2019. Available online: https://dndi.org/press-releases/2019/study-shows-dramatically-shorter-treatment-chagas-effective-and-safer/ (accessed on 21 April 2025).
- Laureano de Souza, M.; Lapierre, T.J.W.J.D.; Marques, G.V.L.; Ferraz, W.R.; Penteado, A.B.; Trossini, G.H.G.; Murta, S.M.F.; de Oliveira, R.B.; Rezende, C.O.; Ferreira, R.S. Molecular targets for Chagas disease: Validation, challenges and lead compounds for widely exploited targets. Expert Opin. Ther. Targets 2023, 27, 911–925. [Google Scholar] [CrossRef] [PubMed]
- Gabaldón-Figueira, J.C.; Martinez-Peinado, N.; Escabia, E.; Ros-Lucas, A.; Chatelain, E.; Scandale, I.; Gascon, J.; Pinazo, M.-J.; Alonso-Padilla, J. State-of-the-art in the drug discovery pathway for Chagas disease: A framework for drug development and target validation. Res. Rep. Trop. Med. 2023, 14, 1–19. [Google Scholar] [CrossRef]
- Gupta, I.; Aggarwal, S.; Singh, K.; Yadav, A.; Khan, S. Ubiquitin proteasome pathway proteins as potential drug targets in parasite Trypanosoma cruzi. Sci. Rep. 2018, 8, 8399. [Google Scholar] [CrossRef]
- Rutherford, K.A.; McManus, K.J. PROTACs: Current and future potential as a precision medicine strategy to combat cancer. Mol. Cancer Ther. 2024, 23, 454–463. [Google Scholar] [CrossRef] [PubMed]
- Martín-Escolano, J.; Marín, C.; Rosales, M.J.; Tsaousis, A.D.; Medina-Carmona, E.; Martín-Escolano, R. An updated view of the Trypanosoma cruzi life cycle: Intervention points for an effective treatment. ACS Infect. Dis. 2022, 8, 1107. [Google Scholar] [CrossRef]
- De Souza, W.; Barrias, E.S. May the epimastigote form of Trypanosoma cruzi be infective? Acta Trop. 2020, 212, 105688. [Google Scholar] [CrossRef]
- Campbell, P.C.; de Graffenried, C.L. Morphogenesis in Trypanosoma cruzi epimastigotes proceeds via a highly asymmetric cell division. PLoS Negl. Trop. Dis. 2023, 17, e0011731. [Google Scholar] [CrossRef] [PubMed]
- Won, M.M.; Krüger, T.; Engstler, M.; Burleigh, B.A. The intracellular amastigote of Trypanosoma cruzi maintains an actively beating flagellum. mBio 2023, 14, e03556-22. [Google Scholar] [CrossRef]
- Arias-del-Angel, J.A.; Santana-Solano, J.; Santillán, M.; Manning-Cela, R.G. Motility patterns of Trypanosoma cruzi trypomastigotes correlate with the efficiency of parasite invasion in vitro. Sci. Rep. 2020, 10, 15894. [Google Scholar] [CrossRef]
- Sousa, A.O.; Gomes, C.; Sá, A.A.; Nascimento, R.J.; Pires, L.L.; Castro, A.M.; Moreno, F.; Teixeira, A.R. The Trypanosoma cruzi kinetoplast DNA minicircle sequences transfer biomarker of the multidrug treatment of Chagas disease. bioRxiv 2021. [Google Scholar] [CrossRef]
- Herreros-Cabello, A.; Callejas-Hernández, F.; Fresno, M.; Gironès, N. Mitochondrial DNA structure in Trypanosoma cruzi. Pathogens 2025, 14, 73. [Google Scholar] [CrossRef]
- Cavalcanti, D.P.; De Souza, W. The kinetoplast of trypanosomatids: From early studies of electron microscopy to recent advances in atomic force microscopy. Scanning 2018, 2018, 9603051. [Google Scholar] [CrossRef]
- Lima, L.; Espinosa-Álvarez, O.; Pinto, C.M.; Cavazzana, M.; Pavan, A.C.; Carranza, J.C.; Lim, B.K.; Campaner, M.; Takata, C.S.A.; Camargo, E.P.; et al. New insights into the evolution of the Trypanosoma cruzi clade provided by a new trypanosome species tightly linked to Neotropical Pteronotus bats and related to an Australian lineage of trypanosomes. Parasites Vectors 2015, 8, 657. [Google Scholar] [CrossRef] [PubMed]
- Gonçalves, C.S.; Ávila, A.R.; De Souza, W.; Motta, M.C.M.; Cavalcanti, D.P. Revisiting the Trypanosoma cruzi metacyclogenesis: Morphological and ultrastructural analyses during cell differentiation. Parasites Vectors 2018, 11, 83. [Google Scholar] [CrossRef] [PubMed]
- Callejas-Hernández, F.; Herreros-Cabello, A.; del Moral-Salmoral, J.; Fresno, M.; Gironès, N. The complete mitochondrial DNA of Trypanosoma cruzi: Maxicircles and minicircles. Front. Cell. Infect. Microbiol. 2021, 11, 672448. [Google Scholar] [CrossRef]
- Gómez-Palacio, A.; Cruz-Saavedra, L.; Van den Broeck, F.; Geerts, M.; Pita, S.; Vallejo, G.A.; Carranza, J.C.; Ramírez, J.D. High-throughput analysis of the Trypanosoma cruzi minicirculome (mcDNA) unveils structural variation and functional diversity. Sci. Rep. 2024, 14, 5578. [Google Scholar] [CrossRef]
- Berna, L.; Greif, G.; Pita, S.; Faral-Tello, P.; Díaz-Viraque, F.; De Cassia Moreira De Souza, R.; Vallejo, G.A.; Alvarez-Valin, F.; Robello, C. Maxicircle architecture and evolutionary insights into Trypanosoma cruzi complex. PLoS Negl. Trop. Dis. 2021, 15, e0009719. [Google Scholar] [CrossRef]
- Povelones, M.L. Beyond replication: Division and segregation of mitochondrial DNA in kinetoplastids. Mol. Biochem. Parasitol. 2014, 196, 53–60. [Google Scholar] [CrossRef]
- Urbina, J.A. Ergosterol biosynthesis and drug development for Chagas disease. Mem. Inst. Oswaldo Cruz 2009, 104, 311–318. [Google Scholar] [CrossRef]
- Lepesheva, G.I.; Waterman, M.R. Sterol 14alpha-demethylase (CYP51) as a therapeutic target for human trypanosomiasis and leishmaniasis. Curr. Top. Med. Chem. 2011, 11, 2060–2071. [Google Scholar] [CrossRef] [PubMed]
- Urbina, J.A.; Docampo, R. Specific chemotherapy of Chagas disease: Controversies and advances. Trends Parasitol. 2003, 19, 495–501. [Google Scholar] [CrossRef]
- Buckner, F.S.; Griffin, J.H.; Wilson, A.J.; Van Voorhis, W.C. Potent anti-Trypanosoma cruzi activities of oxidosqualene cyclase inhibitors. Antimicrob. Agents Chemother. 2001, 45, 1210–1215. [Google Scholar] [CrossRef]
- Lepesheva, G.I.; Park, H.-W.; Hargrove, T.Y.; Vanhollebeke, B.; Wawrzak, Z.; Harp, J.M.; Sundaramoorthy, M.; Nes, W.D.; Pays, E.; Chaudhuri, M.; et al. Crystal structures of Trypanosoma brucei sterol 14α-demethylase and implications for selective treatment of human infections. J. Biol. Chem. 2010, 285, 1773–1780. [Google Scholar] [CrossRef]
- Hargrove, T.Y.; Kim, K.; Correia Soeiro, M.N.; da Silva, C.F.; Batista, D.G.J.; Batista, M.M.; Yazlovitskaya, E.M.; Waterman, M.R.; Sulikowski, G.A.; Lepesheva, G.I. CYP51 structures and structure-based development of novel, pathogen-specific inhibitory scaffolds. Int. J. Parasitol. Drugs Drug Resist. 2012, 2, 178–186. [Google Scholar] [CrossRef]
- Gros, L.; Castillo-Acosta, V.M.; Jiménez, C.J.; Sealey-Cardona, M.; Vargas, S.; Estévez, A.M.; Yardley, V.; Rattray, L.; Croft, S.L.; Ruiz-Perez, L.M.; et al. New azasterols against Trypanosoma brucei: Role of 24-sterol methyltransferase in inhibitor action. Antimicrob. Agents Chemother. 2006, 50, 2595–2601. [Google Scholar] [CrossRef]
- Urbina, J.A.; Vivas, J.; Lazardi, K.; Molina, J.; Payares, G.; Pirns, M.M.; Piras, R. Antiproliferative effects of Δ24(25) sterol methyl transferase inhibitors on Trypanosoma (Schizotrypanum) cruzi: In vitro and in vivo studies. Chemotherapy 1996, 42, 294–307. [Google Scholar] [CrossRef] [PubMed]
- Montalvetti, A.; Bailey, B.N.; Martin, M.B.; Severin, G.W.; Oldfield, E.; Docampo, R. Bisphosphonates are potent inhibitors of Trypanosoma cruzi farnesyl pyrophosphate synthase. J. Biol. Chem. 2001, 276, 33930–33937. [Google Scholar] [CrossRef] [PubMed]
- Szajnman, S.H.; Montalvetti, A.; Wang, Y.; Docampo, R.; Rodriguez, J.B. Bisphosphonates derived from fatty acids are potent inhibitors of Trypanosoma cruzi farnesyl pyrophosphate synthase. Bioorg. Med. Chem. Lett. 2003, 13, 3231–3235. [Google Scholar] [CrossRef]
- Martin, M.B.; Grimley, J.S.; Lewis, J.C.; Heath, H.T.; Bailey, B.N.; Kendrick, H.; Yardley, V.; Caldera, A.; Lira, R.; Urbina, J.A.; et al. Bisphosphonates inhibit the growth of Trypanosoma brucei, Trypanosoma cruzi, Leishmania donovani, Toxoplasma gondii, and Plasmodium falciparum: A potential route to chemotherapy. J. Med. Chem. 2001, 44, 909–916. [Google Scholar] [CrossRef]
- Garzoni, L.R.; Caldera, A.; Meirelles, M.N.L.; de Castro, S.L.; Docampo, R.; Meints, G.A.; Oldfield, E.; Urbina, J.A. Selective in vitro effects of the farnesyl pyrophosphate synthase inhibitor risedronate on Trypanosoma cruzi. Int. J. Antimicrob. Agents 2004, 23, 273–285. [Google Scholar] [CrossRef] [PubMed]
- Urbina, J.A. Ergosterol biosynthesis for the specific treatment of Chagas disease: From basic science to clinical trials. In Trypanosomatid Diseases; Wiley: Hoboken, NJ, USA, 2013; pp. 489–514. [Google Scholar] [CrossRef]
- De Souza, W.; Cola, J.; Rodrigues, F. Sterol biosynthesis pathway as target for anti-trypanosomatid drugs. Interdiscip. Perspect. Infect. Dis. 2009, 2009, 642502. [Google Scholar] [CrossRef] [PubMed]
- Hurtado-Guerrero, R.; Peña-Díaz, J.; Montalvetti, A.; Ruiz-Pérez, L.M.; González-Pacanowska, D. Kinetic properties and inhibition of Trypanosoma cruzi 3-hydroxy-3-methylglutaryl CoA reductase. FEBS Lett. 2002, 510, 141–144. [Google Scholar] [CrossRef] [PubMed]
- Oliveira, L.; Araújo, J.; Costa Júnior, D.; dos Santos Junior, M.; Santos Júnior, A.; Leite, F. Virtual screening for the selection of new candidates to Trypanosoma cruzi farnesyl pyrophosphate synthase inhibitors. J. Braz. Chem. Soc. 2018, 29, 2554–2568. [Google Scholar] [CrossRef]
- Romero-Solano, M.Á.; Rodríguez-Pupo, E.C.; Martinez, I.; Prestegui-Martel, B.; Martínez-Muñoz, A.; Espinoza, B.; Martínez-Otero, D.; López-Guerrero, V.; Esteban Covarrubias, A.K.; Dorazco-González, A. Tetranuclear and dinuclear Cu(II) complexes with risedronate as anti-Trypanosoma cruzi and anti-Leishmania mexicana agents: Synthesis, crystal structures, and biological evaluation. Dalton Trans. 2025, 54, 6043–6059. [Google Scholar] [CrossRef]
- Kamdem, B.P.; Boyom, F.F. Inhibitors of farnesyl diphosphate synthase and squalene synthase: Potential source for anti-Trypanosomatidae drug discovery. Drugs Drug Candidates 2023, 2, 624–652. [Google Scholar] [CrossRef]
- Urbina, J.A.; Concepcion, J.L.; Caldera, A.; Payares, G.; Sanoja, C.; Otomo, T.; Hiyoshi, H. In vitro and in vivo activities of E5700 and ER-119884, two novel orally active squalene synthase inhibitors, against Trypanosoma cruzi. Antimicrob. Agents Chemother. 2004, 48, 2379–2387. [Google Scholar] [CrossRef]
- Rodríguez-Poveda, C.A.; González-Pacanowska, D.; Szajnman, S.H.; Rodríguez, J.B. 2-Alkylaminoethyl-1,1-bisphosphonic acids are potent inhibitors of the enzymatic activity of Trypanosoma cruzi squalene synthase. Antimicrob. Agents Chemother. 2012, 56, 4483–4486. [Google Scholar] [CrossRef]
- Noguera, G.J.; Fabian, L.E.; Lombardo, E.; Finkielsztein, L.M. Studies of 4-arylthiazolylhydrazones derived from 1-indanones as Trypanosoma cruzi squalene epoxidase inhibitors by molecular simulations. Org. Biomol. Chem. 2018, 16, 8525–8536. [Google Scholar] [CrossRef]
- Rabelo, V.W.-H.; Romeiro, N.C.; Abreu, P.A. Design strategies of oxidosqualene cyclase inhibitors: Targeting the sterol biosynthetic pathway. J. Steroid Biochem. Mol. Biol. 2017, 171, 305–317. [Google Scholar] [CrossRef]
- Morand, O.H.; Aebi, J.D.; Dehmlow, H.; Ji, Y.H.; Gains, N.; Lengsfeld, H.; Himber, J. Ro 48-8071, a new 2,3-oxidosqualene:lanosterol cyclase inhibitor lowering plasma cholesterol in hamsters, squirrel monkeys, and minipigs: Comparison to simvastatin. J. Lipid Res. 1997, 38, 373–390. [Google Scholar] [CrossRef] [PubMed]
- Galli, U.; Oliaro-Bosso, S.; Taramino, S.; Venegoni, S.; Pastore, E.; Tron, G.C.; Balliano, G.; Viola, F.; Sorba, G. Design, synthesis, and biological evaluation of new (2E,6E)-10-(dimethylamino)-3,7-dimethyl-2,6-decadien-1-ol ethers as inhibitors of human and Trypanosoma cruzi oxidosqualene cyclase. Bioorg. Med. Chem. Lett. 2007, 17, 220–224. [Google Scholar] [CrossRef]
- Oliaro-Bosso, S.; Viola, F.; Taramino, S.; Tagliapietra, S.; Barge, A.; Cravotto, G.; Balliano, G. Inhibitory effect of umbelliferone aminoalkyl derivatives on oxidosqualene cyclases from S. cerevisiae, T. cruzi, P. carinii, H. sapiens, and A. thaliana: A structure–activity study. ChemMedChem 2007, 2, 226–233. [Google Scholar] [CrossRef] [PubMed]
- Tani, O.; Akutsu, Y.; Ito, S.; Suzuki, T.; Tateishi, Y.; Yamaguchi, T.; Niimi, T.; Namatame, I.; Chiba, Y.; Sakashita, H.; et al. NMR biochemical assay for oxidosqualene cyclase: Evaluation of inhibitor activities on Trypanosoma cruzi and human enzymes. J. Med. Chem. 2018, 61, 5047–5053. [Google Scholar] [CrossRef]
- da Silva Santos-Júnior, P.F.; Schmitt, M.; de Araújo-Júnior, J.X.; da Silva-Júnior, E.F. Sterol 14α-demethylase from trypanosomatid parasites as a promising target for designing new antiparasitic agents. Curr. Top. Med. Chem. 2021, 21, 1900–1921. [Google Scholar] [CrossRef] [PubMed]
- Saccoliti, F.; Madia, V.N.; Tudino, V.; De Leo, A.; Pescatori, L.; Messore, A.; De Vita, D.; Scipione, L.; Brun, R.; Kaiser, M.; et al. Design, synthesis, and biological evaluation of new 1-(aryl-1H-pyrrolyl)(phenyl)methyl-1H-imidazole derivatives as antiprotozoal agents. J. Med. Chem. 2019, 62, 1330–1347. [Google Scholar] [CrossRef]
- Suryadevara, P.K.; Olepu, S.; Lockman, J.W.; Ohkanda, J.; Karimi, M.; Verlinde, C.L.M.J.; Kraus, J.M.; Schoepe, J.; Van Voorhis, W.C.; Hamilton, A.D.; et al. Structurally simple inhibitors of lanosterol 14α-demethylase are efficacious in a rodent model of acute Chagas disease. J. Med. Chem. 2009, 52, 3703–3715. [Google Scholar] [CrossRef]
- Friggeri, L.; Hargrove, T.Y.; Rachakonda, G.; Blobaum, A.L.; Fisher, P.; de Oliveira, G.M.; da Silva, C.F.; Soeiro, M.N.C.; Nes, W.D.; Lindsley, C.W.; et al. Sterol 14α-demethylase structure-based optimization of drug candidates for human infections with the protozoan Trypanosomatidae. J. Med. Chem. 2018, 61, 10910–10921. [Google Scholar] [CrossRef]
- Calvet, C.M.; Vieira, D.F.; Choi, J.Y.; Kellar, D.; Cameron, M.D.; Siqueira-Neto, J.L.; Gut, J.; Johnston, J.B.; Lin, L.; Khan, S.; et al. 4-Aminopyridyl-based CYP51 inhibitors as anti-Trypanosoma cruzi drug leads with improved pharmacokinetic profile and in vivo potency. J. Med. Chem. 2014, 57, 6989–7005. [Google Scholar] [CrossRef]
- Ferreira de Almeida Fiuza, L.; Peres, R.B.; Simões-Silva, M.R.; da Silva, P.B.; Batista, D.G.J.; da Silva, C.F.; da Gama, A.N.S.; Reddy, T.R.K.; Soeiro, M.N.C. Identification of pyrazolo [3,4-e][1,4]thiazepin-based CYP51 inhibitors as potential Chagas disease therapeutic alternative: In vitro and in vivo evaluation, binding mode prediction and SAR exploration. Eur. J. Med. Chem. 2018, 149, 257–268. [Google Scholar] [CrossRef]
- Urbina, J.A.; Vivas, J.; Visbal, G.; Contreras, L.M. Modification of the sterol composition of Trypanosoma cruzi epimastigotes by Δ24(25)-sterol methyl transferase inhibitors and their combinations with ketoconazole. Mol. Biochem. Parasitol. 1995, 73, 199–210. [Google Scholar] [CrossRef] [PubMed]
- Rodrigues, J.C.F.; Attias, M.; Rodriguez, C.; Urbina, J.A.; De Souza, W. Ultrastructural and biochemical alterations induced by 22,26-azasterol, a Δ24(25)-sterol methyltransferase inhibitor, on promastigote and amastigote forms of Leishmania amazonensis. Antimicrob. Agents Chemother. 2002, 46, 487–499. [Google Scholar] [CrossRef]
- Esteva, M.I.; Kettler, K.; Maidana, C.; Fichera, L.; Ruiz, A.M.; Bontempi, E.J.; Andersson, B.; Dahse, H.-M.; Haebel, P.; Ortmann, R.; et al. Benzophenone-based farnesyltransferase inhibitors with high activity against Trypanosoma cruzi. J. Med. Chem. 2005, 48, 7186–7191. [Google Scholar] [CrossRef]
- de Almeida, P.E.; Toledo, D.A.M.; Rodrigues, G.S.C.; D’Avila, H. Lipid bodies as sites of prostaglandin E2 synthesis during Chagas disease: Impact in the parasite escape mechanism. Front. Microbiol. 2018, 9, 499. [Google Scholar] [CrossRef] [PubMed]
- Alves-Ferreira, M.; Carolina, A.; Guimarães, R.; Da Silva, V.P.; Capriles, Z.; Dardenne, L.E.; Degrave, W.M. A New Approach for Potential Drug Target Discovery Through in Silico Metabolic Pathway Analysis Using Trypanosoma cruziGenome Information. GeneDB. Available online: http://www.genedb.org/ (accessed on 4 April 2025).
- Booth, L.A.; Smith, T.K. Lipid metabolism in Trypanosoma cruzi : A review. Mol. Biochem. Parasitol. 2020, 240, 111324. [Google Scholar] [CrossRef]
- Freitas, R.F.; Prokopczyk, I.M.; Zottis, A.; Oliva, G.; Andricopulo, A.D.; Trevisan, M.T.S.; Vilegas, W.; Silva, M.G.V.; Montanari, C.A. Discovery of novel Trypanosoma cruzi glyceraldehyde-3-phosphate dehydrogenase inhibitors. Bioorg. Med. Chem. 2009, 17, 2476–2482. [Google Scholar] [CrossRef]
- Ladame, S.; Castilho, M.S.; Silva, C.H.T.P.; Denier, C.; Hannaert, V.; Périé, J.; Oliva, G.; Willson, M. Crystal structure of Trypanosoma cruzi glyceraldehyde-3-phosphate dehydrogenase complexed with an analogue of 1,3-bisphospho-D-glyceric acid. Eur. J. Biochem. 2003, 270, 4574–4586. [Google Scholar] [CrossRef] [PubMed]
- Rojas-Pirela, M.; Andrade-Alviárez, D.; Rojas, V.; Marcos, M.; Salete-Granado, D.; Chacón-Arnaude, M.; Pérez-Nieto, M.; Kemmerling, U.; Concepción, J.L.; Michels, P.A.M.; et al. Exploring glycolytic enzymes in disease: Potential biomarkers and therapeutic targets in neurodegeneration, cancer and parasitic infections. Open Biol. 2025, 15, 240239. [Google Scholar] [CrossRef]
- Pariona-Llanos, R.; Pavani, R.S.; Reis, M.; Noël, V.; Silber, A.M.; Armelin, H.A.; Cano, M.I.N.; Elias, M.C. Glyceraldehyde 3-phosphate dehydrogenase–telomere association correlates with redox status in Trypanosoma cruzi. PLoS ONE 2015, 10, e0120896. [Google Scholar] [CrossRef]
- Karkowska-Kuleta, J.; Kozik, A. Moonlighting proteins as virulence factors of pathogenic fungi, parasitic protozoa and multicellular parasites. Mol. Oral Microbiol. 2014, 29, 270–283. [Google Scholar] [CrossRef]
- Silva, J.J.N.; Guedes, P.M.M.; Zottis, A.; Balliano, T.L.; Nascimento Silva, F.O.; França Lopes, L.G.; Ellena, J.; Oliva, G.; Andricopulo, A.D.; Franco, D.W.; et al. Novel ruthenium complexes as potential drugs for Chagas disease: Enzyme inhibition and in vitro and in vivo trypanocidal activity. Br. J. Pharmacol. 2010, 160, 260–269. [Google Scholar] [CrossRef] [PubMed]
- Prokopczyk, I.M.; Ribeiro, J.F.; Sartori, G.R.; Sesti-Costa, R.; Silva, J.S.; Freitas, R.F.; Leitão, A.; Montanari, C.A. Integration of methods in cheminformatics and biocalorimetry for the design of trypanosomatid enzyme inhibitors. Future Med. Chem. 2014, 6, 17–33. [Google Scholar] [CrossRef] [PubMed]
- Cordeiro, A.T.; Cáceres, A.J.; Vertommen, D.; Concepción, J.L.; Michels, P.A.M.; Versées, W. The crystal structure of Trypanosoma cruzi glucokinase reveals features determining oligomerization and anomer specificity of hexose-phosphorylating enzymes. J. Mol. Biol. 2007, 372, 1215–1226. [Google Scholar] [CrossRef]
- Racagni, G.E.; Machado de Domenech, E.E. Characterization of Trypanosoma cruzi hexokinase. Mol. Biochem. Parasitol. 1983, 9, 181–188. [Google Scholar] [CrossRef]
- Omolabi, K.F.; Odeniran, P.O.; Olotu, F.A.; Mahmoud, M.E. A mechanistic probe into the dual inhibition of T. cruzi glucokinase and hexokinase in Chagas disease treatment—A stone killing two birds? Chem. Biodivers. 2021, 18, e2000863. [Google Scholar] [CrossRef]
- Sanz-Rodríguez, C.E.; Concepción, J.L.; Pekerar, S.; Oldfield, E.; Urbina, J.A. Bisphosphonates as inhibitors of Trypanosoma cruzi hexokinase: Kinetic and metabolic studies. J. Biol. Chem. 2007, 282, 12377–12387. [Google Scholar] [CrossRef] [PubMed]
- Carey, S.M.; Lander, V.; Gualdrón-López, M.; Michels, P.A.M. At the outer part of the active site in Trypanosoma cruzi glucokinase: The role of phenylalanine 337. Biochimie 2024, 218, 8–19. [Google Scholar] [CrossRef]
- Cáceres, A.J.; Portillo, R.; Acosta, H.; Rosales, D.; Quiñones, W.; Avilan, L.; Salazar, L.; Dubourdieu, M.; Michels, P.A.M.; Concepción, J.L. Molecular and biochemical characterization of hexokinase from Trypanosoma cruzi. Mol. Biochem. Parasitol. 2003, 126, 251–262. [Google Scholar] [CrossRef]
- D’Antonio, E.L.; Deinema, M.S.; Kearns, S.P.; Frey, T.A.; Tanghe, S.; Perry, K.; Roy, T.A.; Gracz, H.S.; Rodriguez, A.; D’Antonio, J. Structure-based approach to the identification of a novel group of selective glucosamine analogue inhibitors of Trypanosoma cruzi glucokinase. Mol. Biochem. Parasitol. 2015, 204, 64–76. [Google Scholar] [CrossRef]
- Mercaldi, G.F.; D’Antonio, E.L.; Aguessi, A.; Rodriguez, A.; Cordeiro, A.T. Discovery of antichagasic inhibitors by high-throughput screening with Trypanosoma cruzi glucokinase. Bioorg. Med. Chem. Lett. 2019, 29, 1948–1953. [Google Scholar] [CrossRef]
- Carey, S.M.; O’Neill, D.M.; Conner, G.B.; Sherman, J.; Rodriguez, A.; D’Antonio, E.L. Discovery of strong 3-nitro-2-phenyl-2H-chromene analogues as antitrypanosomal agents and inhibitors of Trypanosoma cruzi glucokinase. Int. J. Mol. Sci. 2024, 25, 4319. [Google Scholar] [CrossRef] [PubMed]
- Fernandes, P.M.; Kinkead, J.; McNae, I.W.; Vásquez-Valdivieso, M.A.; Wear, M.A.; Michels, P.A.M.; Walkinshaw, M.D. Kinetic and structural studies of Trypanosoma and Leishmania phosphofructokinases show evolutionary divergence and identify AMP as a switch regulating glycolysis versus gluconeogenesis. FEBS J. 2020, 287, 2847. [Google Scholar] [CrossRef] [PubMed]
- Rodríguez, E.; Lander, N.; Ramirez, J.L. Molecular and biochemical characterisation of Trypanosoma cruzi phosphofructokinase. Mem. Inst. Oswaldo Cruz 2009, 104, 745–748. [Google Scholar] [CrossRef] [PubMed]
- Brimacombe, K.R.; Walsh, M.J.; Liu, L.; Vásquez-Valdivieso, M.G.; Morgan, H.P.; McNae, I.; Fothergill-Gilmore, L.A.; Michels, P.A.M.; Auld, D.S.; Simeonov, A.; et al. Identification of ML251, a potent inhibitor of T. brucei and T. cruzi phosphofructokinase. ACS Med. Chem. Lett. 2014, 5, 12–17. [Google Scholar] [CrossRef]
- Esteve, M.I.; Cazzulo, J.J. The 6-phosphogluconate dehydrogenase from Trypanosoma cruzi: The absence of two inter-subunit salt bridges as a reason for enzyme instability. Mol. Biochem. Parasitol. 2004, 133, 197–207. [Google Scholar] [CrossRef]
- Naciuk, F.F.; Faria, J.D.N.; Eufrásio, A.G.; Cordeiro, A.T.; Bruder, M. Development of selective steroid inhibitors for the glucose-6-phosphate dehydrogenase from Trypanosoma cruzi. ACS Med. Chem. Lett. 2020, 11, 1250–1256. [Google Scholar] [CrossRef]
- Dardonville, C.; Rinaldi, E.; Barrett, M.P.; Brun, R.; Gilbert, I.H.; Hanau, S. Selective inhibition of Trypanosoma brucei 6-phosphogluconate dehydrogenase by high-energy intermediate and transition-state analogues. J. Med. Chem. 2004, 47, 3427–3437. [Google Scholar] [CrossRef]
- Glockzin, K.; Kostomiris, D.; Minnow, Y.V.T.; Suthagar, K.; Clinch, K.; Gai, S.; Buckler, J.N.; Schramm, V.L.; Tyler, P.C.; Meek, T.D.; et al. Kinetic characterization and inhibition of Trypanosoma cruzi hypoxanthine-guanine phosphoribosyltransferases. Biochemistry 2022, 61, 2088–2105. [Google Scholar] [CrossRef]
- Vidhya, V.M.; Ponnuraj, K. Structure-based virtual screening and computational study towards identification of novel inhibitors of hypoxanthine-guanine phosphoribosyltransferase of Trypanosoma cruzi. J. Cell. Biochem. 2021, 122, 1701–1714. [Google Scholar] [CrossRef]
- Glockzin, K.; Meneely, K.M.; Hughes, R.; Maatouk, S.W.; Piña, G.E.; Suthagar, K.; Clinch, K.; Buckler, J.N.; Lamb, A.L.; Tyler, P.C.; et al. Kinetic and Structural Characterization of Trypanosoma cruzi Hypoxanthine-Guanine-Xanthine Phosphoribosyltransferases and Repurposing of Transition-State Analogue Inhibitors. Biochemistry 2023, 62, 2182–2201. [Google Scholar] [CrossRef]
- Reche, P.; Arrebola, R.; Santi, D.V.; Gonzalez-Pacanowska, D.; Ruiz-Perez, L.M. Expression and characterization of the Trypanosoma cruzi dihydrofolate reductase domain. Mol. Biochem. Parasitol. 1996, 76, 175–185. [Google Scholar] [CrossRef] [PubMed]
- Senkovich, O.; Schormann, N.; Chattopadhyay, D. Structures of dihydrofolate reductase-thymidylate synthase of Trypanosoma cruzi in the folate-free state and in complex with two antifolate drugs, trimetrexate and methotrexate. Acta Crystallogr. D Biol. Crystallogr. 2009, 65, 704–716. [Google Scholar] [CrossRef] [PubMed]
- Schormann, N.; Senkovich, O.; Walker, K.; Wright, D.L.; Anderson, A.C.; Rosowsky, A.; Ananthan, S.; Shinkre, B.; Velu, S.; Chattopadhyay, D. Structure-based approach to pharmacophore identification, in silico screening, and three-dimensional quantitative structure–activity relationship studies for inhibitors of Trypanosoma cruzi dihydrofolate reductase function. Proteins 2008, 73, 889–901. [Google Scholar] [CrossRef] [PubMed]
- Schormann, N.; Velu, S.E.; Murugesan, S.; Senkovich, O.; Walker, K.; Chenna, B.C.; Shinkre, B.; Desai, A.; Chattopadhyay, D. Synthesis and characterization of potent inhibitors of Trypanosoma cruzi dihydrofolate reductase. Bioorg. Med. Chem. 2010, 18, 4056–4066. [Google Scholar] [CrossRef] [PubMed]
- Robello, C.; Navarro, P.; Castanys, S.; Gamarro, F. A pteridine reductase gene ptr1 contiguous to a P-glycoprotein confers resistance to antifolates in Trypanosoma cruzi. Mol. Biochem. Parasitol. 1997, 90, 525–535. [Google Scholar] [CrossRef]
- Panecka-Hofman, J.; Poehner, I. Structure and dynamics of pteridine reductase 1: The key phenomena relevant to enzyme function and drug design. Eur. Biophys. J. 2023, 52, 521. [Google Scholar] [CrossRef]
- Schormann, N.; Pal, B.; Senkovich, O.; Carson, M.; Howard, A.; Smith, C.; DeLucas, L.; Chattopadhyay, D. Crystal structure of Trypanosoma cruzi pteridine reductase 2 in complex with a substrate and an inhibitor. J. Struct. Biol. 2005, 152, 64–75. [Google Scholar] [CrossRef]
- Cavazzuti, A.; Paglietti, G.; Hunter, W.N.; Gamarro, F.; Piras, S.; Loriga, M.; Alleca, S.; Corona, P.; McLuskey, K.; Tulloch, L.; et al. Discovery of potent pteridine reductase inhibitors to guide antiparasite drug development. Proc. Natl. Acad. Sci. USA 2008, 105, 1448–1453. [Google Scholar] [CrossRef]
- Tiwari, K.; Dubey, V.K. Fresh insights into the pyrimidine metabolism in the trypanosomatids. Parasit. Vectors 2018, 11, 87. [Google Scholar] [CrossRef]
- Hashimoto, M.; Morales, J.; Fukai, Y.; Suzuki, S.; Takamiya, S.; Tsubouchi, A.; Inoue, S.; Inoue, M.; Kita, K.; Harada, S.; et al. Critical importance of the De Novo pyrimidine biosynthesis pathway for Trypanosoma cruzi growth in the mammalian host cell cytoplasm. Biochem. Biophys. Res. Commun. 2012, 417, 1002–1006. [Google Scholar] [CrossRef]
- Inaoka, D.K.; Iida, M.; Hashimoto, S.; Tabuchi, T.; Kuranaga, T.; Balogun, E.O.; Honma, T.; Tanaka, A.; Harada, S.; Nara, T.; et al. Design and synthesis of potent substrate-based inhibitors of the Trypanosoma cruzi dihydroorotate dehydrogenase. Bioorg. Med. Chem. 2017, 25, 1465–1470. [Google Scholar] [CrossRef] [PubMed]
- de Jesus, G.C.; Ribeiro, T.S.; Andricopulo, A.D.; Ferreira, L.L.G. Molecular docking and quantitative structure–activity relationships for a series of Trypanosoma cruzi dihydroorotate dehydrogenase inhibitors. Braz. J. Pharm. Sci. 2025, 61, e24064. [Google Scholar] [CrossRef]
- González-Montero, M.C.; Andrés-Rodríguez, J.; García-Fernández, N.; Pérez-Pertejo, Y.; Reguera, R.M.; Balaña-Fouce, R.; García-Estrada, C. Targeting trypanothione metabolism in trypanosomatids. Molecules 2024, 29, 2214. [Google Scholar] [CrossRef]
- Jockers-Scherübl, M.C.; Schirmer, R.H.; Krauth-Siegel, R.L. Trypanothione reductase from Trypanosoma cruzi: Catalytic properties of the enzyme and inhibition studies with trypanocidal compounds. Eur. J. Biochem. 1989, 180, 267–272. [Google Scholar] [CrossRef] [PubMed]
- Piacenza, L.; Peluffo, G.; Alvarez, M.N.; Martínez, A.; Radi, R. Trypanosoma cruzi antioxidant enzymes as virulence factors in Chagas disease. Antioxid. Redox Signal. 2013, 19, 723. [Google Scholar] [CrossRef] [PubMed]
- Beltran-Hortelano, I.; Perez-Silanes, S.; Galiano, S. Trypanothione reductase and superoxide dismutase as current drug targets for Trypanosoma cruzi: An overview of compounds with activity against Chagas disease. Curr. Med. Chem. 2017, 24, 1066–1138. [Google Scholar] [CrossRef]
- Baillet, S.; Buisine, E.; Horvath, D.; Maes, L.; Bonnet, B.; Sergheraert, C. 2-Amino diphenylsulfides as inhibitors of trypanothione reductase: Modification of the side chain. Bioorg. Med. Chem. 1996, 4, 891–899. [Google Scholar] [CrossRef]
- Persch, E.; Bryson, S.; Todoroff, N.K.; Eberle, C.; Thelemann, J.; Dirdjaja, N.; Kaiser, M.; Weber, M.; Derbani, H.; Brun, R.; et al. Binding to large enzyme pockets: Small-molecule inhibitors of trypanothione reductase. ChemMedChem 2014, 9, 1880–1891. [Google Scholar] [CrossRef]
- Eberle, C.; Lauber, B.S.; Fankhauser, D.; Kaiser, M.; Brun, R.; Krauth-Siegel, R.L.; Diederich, F. Improved inhibitors of trypanothione reductase by combination of motifs: Synthesis, inhibitory potency, binding mode, and antiprotozoal activities. ChemMedChem 2011, 6, 292–301. [Google Scholar] [CrossRef]
- Meiering, S.; Inhoff, O.; Mies, J.; Vincek, A.; Garcia, G.; Kramer, B.; Dormeyer, M.; Krauth-Siegel, R.L. Inhibitors of Trypanosoma cruzi trypanothione reductase revealed by virtual screening and parallel synthesis. J. Med. Chem. 2005, 48, 4793–4802. [Google Scholar] [CrossRef]
- Mendonça, A.A.S.; Coelho, C.M.; Veloso, M.P.; Caldas, I.S.; Gonçalves, R.V.; Teixeira, A.L.; de Miranda, A.S.; Novaes, R.D. Relevance of trypanothione reductase inhibitors on Trypanosoma cruzi infection: A systematic review, meta-analysis, and in silico integrated approach. Oxid. Med. Cell. Longev. 2018, 2018, 8676578. [Google Scholar] [CrossRef] [PubMed]
- Mesías, A.C.; Sasoni, N.; Arias, D.G.; Pérez Brandán, C.; Orban, O.C.F.; Kunick, C.; Robello, C.; Comini, M.A.; Garg, N.J.; Zago, M.P. Trypanothione synthetase confers growth, survival advantage and resistance to anti-protozoal drugs in Trypanosoma cruzi. Free Radic. Biol. Med. 2019, 130, 23–34. [Google Scholar] [CrossRef]
- Gomez, J.; Coll, M.; Guarise, C.; Cifuente, D.; Masone, D.; Tello, P.F.; Piñeyro, M.D.; Robello, C.; Reta, G.; Sosa, M.Á.; et al. New insights into the pro-oxidant mechanism of dehydroleucodine on Trypanosoma cruzi. Sci. Rep. 2024, 14, 18875. [Google Scholar] [CrossRef] [PubMed]
- Benítez, D.; Medeiros, A.; Fiestas, L.; Panozzo-Zenere, E.A.; Maiwald, F.; Prousis, K.C.; Roussaki, M.; Calogeropoulou, T.; Detsi, A.; Jaeger, T.; et al. Identification of novel chemical scaffolds inhibiting trypanothione synthetase from pathogenic trypanosomatids. PLoS Negl. Trop. Dis. 2016, 10, e0004617. [Google Scholar] [CrossRef]
- Peloso, E.F.; Dias, L.; Queiroz, R.M.L.; Leme, A.F.P.P.; Pereira, C.N.; Carnielli, C.M.; Werneck, C.C.; Sousa, M.V.; Ricart, C.A.O.; Gadelha, F.R. Trypanosoma cruzi mitochondrial tryparedoxin peroxidase is located throughout the cell and its pull-down provides one step towards the understanding of its mechanism of action. Biochim. Biophys. Acta Proteins Proteom. 2016, 1864, 1–10. [Google Scholar] [CrossRef]
- López, L.; Chiribao, M.L.; Girard, M.C.; Gómez, K.A.; Carasi, P.; Fernandez, M.; Hernandez, Y.; Robello, C.; Freire, T.; Piñeyro, M.D. The cytosolic tryparedoxin peroxidase from Trypanosoma cruzi induces a pro-inflammatory Th1 immune response in a peroxidatic cysteine-dependent manner. Immunology 2021, 163, 46–59. [Google Scholar] [CrossRef] [PubMed]
- Kohatsu, A.A.N.; Silva, F.A.J.; Francisco, A.I.; Rimoldi, A.; Silva, M.T.A.; Vargas, M.D.; da Rosa, J.A.; Cicarelli, R.M.B. Differential expression on mitochondrial tryparedoxin peroxidase (mTcTXNPx) in Trypanosoma cruzi after ferrocenyl diamine hydrochlorides treatments. Braz. J. Infect. Dis. 2017, 21, 125–132. [Google Scholar] [CrossRef]
- Merritt, C.; Silva, L.E.; Tanner, A.L.; Stuart, K.; Pollastri, M.P. Kinases as druggable targets in trypanosomatid protozoan parasites. Chem. Rev. 2014, 114, 11280–11304. [Google Scholar] [CrossRef]
- Bustamante, C.; Díez-Mejía, A.F.; Arbeláez, N.; Soares, M.J.; Robledo, S.M.; Ochoa, R.; Varela-M., R.E.; Marín-Villa, M. In Silico, In Vitro, and Pharmacokinetic Studies of UBMC-4, a Potential Novel Compound for Treating against Trypanosoma cruzi. Pathogens 2022, 11, 616. [Google Scholar] [CrossRef]
- Stadler, K.A.; Ortiz-Joya, L.J.; Singh Sahrawat, A.; Buhlheller, C.; Gruber, K.; Pavkov-Keller, T.; O’Hagan, T.B.; Guarné, A.; Pulido, S.; Marín-Villa, M.; et al. Structural investigation of Trypanosoma cruzi Akt-like kinase as drug target against Chagas disease. Sci. Rep. 2024, 14, 59654. [Google Scholar] [CrossRef]
- Chiurillo, M.A.; Jensen, B.C.; Docampo, R. Drug target validation of the protein kinase AEK1, essential for proliferation, host cell invasion, and intracellular replication of the human pathogen Trypanosoma cruzi. Microbiol. Spectr. 2021, 9, e00738-21. [Google Scholar] [CrossRef] [PubMed]
- Graefe, S.E.B.; Wiesgigl, M.; Gaworski, I.; Macdonald, A.; Clos, J. Inhibition of HSP90 in Trypanosoma cruzi induces a stress response but no stage differentiation. Eukaryot. Cell 2002, 1, 936–943. [Google Scholar] [CrossRef] [PubMed]
- Martínez-Peinado, N.; Martori, C.; Cortes-Serra, N.; Sherman, J.; Rodríguez, A.; Gascón, J.; Alberola, J.; Pinazo, M.-J.; Rodriguez-Cortes, A.; Alonso-Padilla, J. Anti-Trypanosoma cruzi activity of metabolism modifier compounds. Int. J. Mol. Sci. 2021, 22, 688. [Google Scholar] [CrossRef]
- Costa-Silva, H.M.; Resende, B.C.; Umaki, A.C.S.; Prado, W.; da Silva, M.S.; Virgílio, S.; Macedo, A.M.; Pena, S.D.J.; Tahara, E.B.; Tosi, L.R.O.; et al. DNA topoisomerase 3α is involved in homologous recombination repair and replication stress response in Trypanosoma cruzi. Front. Cell Dev. Biol. 2021, 9, 633195. [Google Scholar] [CrossRef]
- Jobe, M.; Anwuzia-Iwegbu, C.; Banful, A.; Bosier, E.; Iqbal, M.; Jones, K.; Lecutier, S.J.; Lepper, K.; Redmond, M.; Ross-Parker, A.; et al. Differential in vitro activity of the DNA topoisomerase inhibitor idarubicin against Trypanosoma rangeli and Trypanosoma cruzi. Mem. Inst. Oswaldo Cruz 2012, 107, 946–950. [Google Scholar] [CrossRef] [PubMed]
- Rao, S.P.S.; Gould, M.K.; Noeske, J.; Saldivia, M.; Jumani, R.S.; Ng, P.S.; René, O.; Chen, Y.L.; Kaiser, M.; Ritchie, R.; et al. Cyanotriazoles are selective topoisomerase II poisons that rapidly cure trypanosome infections. Science 2023, 380, 1349–1356. [Google Scholar] [CrossRef]
- Alonso, V.L.; Tavernelli, L.E.; Pezza, A.; Cribb, P.; Ritagliati, C.; Serra, E. Aim for the readers! Bromodomains as new targets against Chagas’ disease. Curr. Med. Chem. 2018, 26, 6544–6563. [Google Scholar] [CrossRef]
- Alonso, V.L.; Ritagliati, C.; Cribb, P.; Cricco, J.A.; Serra, E.C. Overexpression of bromodomain factor 3 in Trypanosoma cruzi (TcBDF3) affects differentiation of the parasite and protects it against bromodomain inhibitors. FEBS J. 2016, 283, 2051–2066. [Google Scholar] [CrossRef]
- Laurin, C.M.C.; Bluck, J.P.; Chan, A.K.N.; Keller, M.; Boczek, A.; Scorah, A.R.; See, K.F.L.; Jennings, L.E.; Hewings, D.S.; Woodhouse, F.; et al. Fragment-based identification of ligands for bromodomain-containing factor 3 of Trypanosoma cruzi. ACS Infect. Dis. 2021, 7, 2238–2249. [Google Scholar] [CrossRef]
- Alonso, V.L.; Escalante, A.M.; Rodríguez Araya, E.; Frattini, G.; Tavernelli, L.E.; Moreno, D.M.; Furlan, R.L.E.; Serra, E. 1,3,4-Oxadiazoles as inhibitors of the atypical member of the BET family bromodomain factor 3 from Trypanosoma cruzi (TcBDF3). Front. Microbiol. 2024, 15, 1465672. [Google Scholar] [CrossRef]
- Buschiazzo, A.; Amaya, M.F.; Cremona, M.L.; Frasch, A.C.; Alzari, P.M. The crystal structure and mode of action of trans-sialidase, a key enzyme in Trypanosoma cruzi pathogenesis. Mol. Cell 2002, 10, 757–768. [Google Scholar] [CrossRef] [PubMed]
- Nardy, A.F.F.R.; Freire-de-Lima, C.G.; Pérez, A.R.; Morrot, A. Role of Trypanosoma cruzi trans-sialidase on the escape from host immune surveillance. Front. Microbiol. 2016, 7, 348. [Google Scholar] [CrossRef] [PubMed]
- Buschiazzo, A.; Muiá, R.; Larrieux, N.; Pitcovsky, T.; Mucci, J.; Campetella, O. Trypanosoma cruzi trans-sialidase in complex with a neutralizing antibody: Structure/function studies towards the rational design of inhibitors. PLoS Pathog. 2012, 8, e1002474. [Google Scholar] [CrossRef]
- Neres, J.; Brewer, M.L.; Ratier, L.; Botti, H.; Buschiazzo, A.; Edwards, P.N.; Mortenson, P.N.; Charlton, M.H.; Alzari, P.M.; Frasch, A.C.; et al. Discovery of novel inhibitors of Trypanosoma cruzi trans-sialidase from in silico screening. Bioorg. Med. Chem. Lett. 2009, 19, 589–596. [Google Scholar] [CrossRef]
- Kashif, M.; Chacón-Vargas, K.F.; López-Cedillo, J.C.; Nogueda-Torres, B.; Paz-González, A.D.; Ramírez-Moreno, E.; Agusti, R.; Uhrig, M.L.; Reyes-Arellano, A.; Peralta-Cruz, J.; et al. Synthesis, molecular docking and biological evaluation of novel phthaloyl derivatives of 3-amino-3-aryl propionic acids as inhibitors of Trypanosoma cruzi trans-sialidase. Eur. J. Med. Chem. 2018, 156, 252–268. [Google Scholar] [CrossRef]
- Vázquez-Jiménez, L.K.; Paz-González, A.D.; Juárez-Saldivar, A.; Uhrig, M.L.; Agusti, R.; Reyes-Arellano, A.; Nogueda-Torres, B.; Rivera, G. Structure-based virtual screening of new benzoic acid derivatives as Trypanosoma cruzi trans-sialidase inhibitors. Med. Chem. 2021, 17, 724–731. [Google Scholar] [CrossRef]
- Santos, V.C.; Leite, P.G.; Santos, L.H.; Pascutti, P.G.; Kolb, P.; Machado, F.S.; Ferreira, R.S. Structure-based discovery of novel cruzain inhibitors with distinct trypanocidal activity profiles. Eur. J. Med. Chem. 2023, 257, 115498. [Google Scholar] [CrossRef]
- McGrath, M.E.; Eakin, A.E.; Engel, J.C.; McKerrow, J.H.; Craik, C.S.; Fletterick, R.J. The crystal structure of cruzain: A therapeutic target for Chagas’ disease. J. Mol. Biol. 1995, 247, 251–259. [Google Scholar] [CrossRef] [PubMed]
- Müller, P.; Meta, M.; Meidner, J.L.; Schwickert, M.; Meyr, J.; Schwickert, K.; Kersten, C.; Zimmer, C.; Hammerschmidt, S.J.; Frey, A.; et al. Investigation of the compatibility between warheads and peptidomimetic sequences of protease inhibitors—A comprehensive reactivity and selectivity study. Int. J. Mol. Sci. 2023, 24, 7226. [Google Scholar] [CrossRef]
- Silva, J.R.A.; Cianni, L.; Araujo, D.; Batista, P.H.J.; De Vita, D.; Rosini, F.; Leitão, A.; Lameira, J.; Montanari, C.A. Assessment of the cruzain cysteine protease reversible and irreversible covalent inhibition mechanism. J. Chem. Inf. Model. 2020, 60, 1666–1677. [Google Scholar] [CrossRef]
- Brak, K.; Doyle, P.S.; McKerrow, J.H.; Ellman, J.A. Identification of a new class of nonpeptidic inhibitors of cruzain. J. Am. Chem. Soc. 2008, 130, 6404. [Google Scholar] [CrossRef] [PubMed]
- Vital, D.G.; Damasceno, F.S.; Rapado, L.N.; Silber, A.M.; Vilella, F.S.; Ferreira, R.S.; Maltarollo, V.G.; Trossini, G.H.G. Application of bioisosterism in design of the semicarbazone derivatives as cruzain inhibitors: A theoretical and experimental study. J. Biomol. Struct. Dyn. 2017, 35, 1244–1259. [Google Scholar] [CrossRef] [PubMed]
- Rosas-Jimenez, J.G.; Garcia-Revilla, M.A.; Madariaga-Mazon, A.; Martinez-Mayorga, K. Predictive global models of cruzain inhibitors with large chemical coverage. ACS Omega 2021, 6, 6722–6735. [Google Scholar] [CrossRef] [PubMed]
- Gonçalves, R.B.; Ferraz, W.R.; Calil, R.L.; Scotti, M.T.; Trossini, G.H.G. Convergent QSAR models for the prediction of cruzain inhibitors. ACS Omega 2023, 8, 38961–38982. [Google Scholar] [CrossRef]
- Prates, J.L.B.; Lopes, J.R.; Chin, C.M.; Ferreira, E.I.; dos Santos, J.L.; Scarim, C.B. Discovery of novel inhibitors of cruzain cysteine protease of Trypanosoma cruzi. Curr. Med. Chem. 2024, 31, 2285–2308. [Google Scholar] [CrossRef]
- Barbosa Da Silva, E.; Sharma, V.; Hernandez-Alvarez, L.; Tang, A.H.; Stoye, A.; O’Donoghue, A.J.; Gerwick, W.H.; Payne, R.J.; McKerrow, J.H.; Podust, L.M. Intramolecular interactions enhance the potency of gallinamide A analogues against Trypanosoma cruzi. J. Med. Chem. 2022, 65, 4255–4269. [Google Scholar] [CrossRef]
- de Souza, M.L.; de Oliveira Rezende Junior, C.; Ferreira, R.S.; Espinoza Chávez, R.M.; Ferreira, L.L.G.; Slafer, B.W.; Magalhães, L.G.; Krogh, R.; Oliva, G.; Cruz, F.C.; et al. Discovery of potent, reversible, and competitive cruzain inhibitors with trypanocidal activity: A structure-based drug design approach. J. Chem. Inf. Model. 2020, 60, 1028–1041. [Google Scholar] [CrossRef]
- Cardoso, M.V.O.; de Siqueira, L.R.P.; da Silva, E.B.; Costa, L.B.; Hernandes, M.Z.; Rabello, M.M.; Ferreira, R.S.; da Cruz, L.F.; Moreira, D.R.M.; Pereira, V.R.A.; et al. 2-Pyridyl thiazoles as novel anti-Trypanosoma cruzi agents: Structural design, synthesis and pharmacological evaluation. Eur. J. Med. Chem. 2014, 86, 48–59. [Google Scholar] [CrossRef]
- Hall, B.S.; Wilkinson, S.R. Activation of benznidazole by trypanosomal type I nitroreductases results in glyoxal formation. Antimicrob. Agents Chemother. 2012, 56, 115. [Google Scholar] [CrossRef]
- Cirqueira, M.L.; Bortot, L.O.; Bolean, M.; Aleixo, M.A.A.; Luccas, P.H.; Costa-Filho, A.J.; Ramos, A.P.; Ciancaglini, P.; Nonato, M.C. Trypanosoma cruzi nitroreductase: Structural features and interaction with biological membranes. Int. J. Biol. Macromol. 2022, 221, 891–899. [Google Scholar] [CrossRef]
- Trochine, A.; Creek, D.J.; Faral-Tello, P.; Barrett, M.P.; Robello, C. Benznidazole biotransformation and multiple targets in Trypanosoma cruzi revealed by metabolomics. PLoS Negl. Trop. Dis. 2014, 8, e2844. [Google Scholar] [CrossRef] [PubMed]
- Mejía-Jaramillo, A.M.; Fernández, G.J.; Palacio, L.; Triana-Chávez, O. Gene expression study using real-time PCR identifies an NTR gene as a major marker of resistance to benznidazole in Trypanosoma cruzi. Parasites Vectors 2011, 4, 169. [Google Scholar] [CrossRef]
- Menozzi, C.A.C.; França, R.R.F.; Luccas, P.H.; Baptista, M.S.; Fernandes, T.V.A.; Hoelz, L.V.B.; Sales Junior, P.A.; Murta, S.M.F.; Romanha, A.; Galvão, B.V.D.; et al. Anti-Trypanosoma cruzi activity, mutagenicity, hepatocytotoxicity and nitroreductase enzyme evaluation of 3-nitrotriazole, 2-nitroimidazole and triazole derivatives. Molecules 2023, 28, 7461. [Google Scholar] [CrossRef] [PubMed]
- Carvalho, D.B.; Costa, P.A.N.; Portapilla, G.B.; das Neves, A.R.; Shiguemoto, C.Y.K.; Pelizaro, B.I.; Silva, F.; Piranda, E.M.; Arruda, C.C.P.; Gaspari, P.D.M.; et al. Design, synthesis and antitrypanosomatid activity of 2-nitroimidazole-3,5-disubstituted isoxazole compounds based on benznidazole. Eur. J. Med. Chem. 2023, 260, 115451. [Google Scholar] [CrossRef]
- Pitombeira, M.C.S.R.; Júnior, P.A.S.; Murta, S.M.F.; Romanha, A.; Luccas, P.H.; Nonato, M.C.; Rocha, R.E.O.; Ferreira, R.S.; da Silveira, F.F.; Castelo-Branco, F.S.; et al. New 2-nitroimidazole-N-acylhydrazones, analogs of benznidazole, as anti-Trypanosoma cruzi agents. Arch. Pharm. 2024, 357, e2400059. [Google Scholar] [CrossRef]
- Zhong, G.; Chang, X.; Xie, W.; Zhou, X. Targeted protein degradation: Advances in drug discovery and clinical practice. Signal Transduct. Target. Ther. 2024, 9, 308. [Google Scholar] [CrossRef]
- Pettersson, M.; Crews, C.M. PROteolysis TArgeting Chimeras (PROTACs)—Past, present and future. Drug Discov. Today Technol. 2019, 31, 15–27. [Google Scholar] [CrossRef]
- Liu, Z.; Hu, M.; Yang, Y.; Du, C.; Zhou, H.; Liu, C.; Chen, Y.; Fan, L.; Ma, H.; Gong, Y.; et al. An overview of PROTACs: A promising drug discovery paradigm. Mol. Biomed. 2022, 3, 46. [Google Scholar] [CrossRef]
- Békés, M.; Langley, D.R.; Crews, C.M. PROTAC targeted protein degraders: The past is prologue. Nat. Rev. Drug Discov. 2022, 21, 181–200. [Google Scholar] [CrossRef]
- Zhao, L.; Zhao, J.; Zhong, K.; Tong, A.; Jia, D. Targeted protein degradation: Mechanisms, strategies and application. Signal Transduct. Target. Ther. 2022, 7, 113. [Google Scholar] [CrossRef]
- Ciulli, A.; Trainor, N. A beginner’s guide to PROTACs and targeted protein degradation. Biochemist 2021, 43, 74–79. [Google Scholar] [CrossRef]
- Graham, H. The mechanism of action and clinical value of PROTACs: A graphical review. Cell Signal 2022, 99, 110446. [Google Scholar] [CrossRef] [PubMed]
- Moreau, K.; Coen, M.; Zhang, A.X.; Pachl, F.; Castaldi, M.P.; Dahl, G.; Boyd, H.; Scott, C.; Newham, P. Proteolysis-targeting chimeras in drug development: A safety perspective. Br. J. Pharmacol. 2020, 177, 1709–1718. [Google Scholar] [CrossRef] [PubMed]
- Wang, C.; Zhang, Y.; Zhang, T.; Shi, L.; Geng, Z.; Xing, D. Proteolysis-targeting chimaeras (PROTACs) as pharmacological tools and therapeutic agents: Advances and future challenges. J. Enzym. Inhib. Med. Chem. 2022, 37, 1667–1693. [Google Scholar] [CrossRef]
- Wang, C.; Zhang, Y.; Chen, W.; Wu, Y.; Xing, D. New-generation advanced PROTACs as potential therapeutic agents in cancer therapy. Mol. Cancer 2024, 23, 110. [Google Scholar] [CrossRef]
- Zhou, C.; Yang, S.; Wang, J.; Pan, W.; Yao, H.; Li, G.; Niu, M. Recent advances in PROTAC-based antiviral and antibacterial therapeutics. Bioorganic Chem. 2025, 160, 108437. [Google Scholar] [CrossRef]
- Danazumi, A.U.; Ishmam, I.T.; Idris, S.; Izert, M.A.; Balogun, E.O.; Górna, M.W. Targeted protein degradation might present a novel therapeutic approach in the fight against African trypanosomiasis. Eur. J. Pharm. Sci. 2023, 186, 106451. [Google Scholar] [CrossRef]
- Li, R.; Liu, M.; Yang, Z.; Li, J.; Gao, Y.; Tan, R. Proteolysis-targeting chimeras (PROTACs) in cancer therapy: Present and future. Molecules 2022, 27, 8828. [Google Scholar] [CrossRef]
- Zengerle, M.; Chan, K.H.; Ciulli, A. Selective small molecule induced degradation of the BET bromodomain protein BRD4. ACS Chem. Biol. 2015, 10, 1770–1777. [Google Scholar] [CrossRef]
- Qi, S.M.; Dong, J.; Xu, Z.Y.; Cheng, X.D.; Zhang, W.D.; Qin, J.J. PROTAC: An effective targeted protein degradation strategy for cancer therapy. Front. Pharmacol. 2021, 12, 692574. [Google Scholar] [CrossRef]
- Churcher, I. PROTAC-induced protein degradation in drug discovery: Breaking the rules or just making new ones? J. Med. Chem. 2018, 61, 444–452. [Google Scholar] [CrossRef] [PubMed]
- Piya, S.; Mu, H.; Bhattacharya, S.; Lorenzi, P.L.; Davis, R.E.; McQueen, T.; Ruvolo, V.; Baran, N.; Wang, Z.; Qian, Y.; et al. BETP degradation simultaneously targets acute myelogenous leukemic stem cells and the microenvironment. J. Clin. Investig. 2019, 129, 1878–1894. [Google Scholar] [CrossRef]
- Swain, S.M.; Shastry, M.; Hamilton, E. Targeting HER2-positive breast cancer: Advances and future directions. Nat. Rev. Drug Discov. 2022, 22, 101–126. [Google Scholar] [CrossRef]
- Maneiro, M.; Forte, N.; Shchepinova, M.M.; Kounde, C.S.; Chudasama, V.; Baker, J.R.; Tate, E.W. Antibody–PROTAC conjugates enable HER2-dependent targeted protein degradation of BRD4. ACS Chem. Biol. 2020, 15, 1306–1312. [Google Scholar] [CrossRef]
- Gopalsamy, A. Selectivity through targeted protein degradation (TPD). J. Med. Chem. 2022, 65, 8113–8126. [Google Scholar] [CrossRef] [PubMed]
- Morreale, F.E.; Kleine, S.; Leodolter, J.; Junker, S.; Hoi, D.M.; Ovchinnikov, S.; Okun, A.; Kley, J.; Kurzbauer, R.; Junk, L.; et al. BacPROTACs mediate targeted protein degradation in bacteria. Cell 2022, 185, 2338–2353.e18. [Google Scholar] [CrossRef]
- Sarathy, J.P.; Aldrich, C.C.; Go, M.L.; Dick, T. PROTAC antibiotics: The time is now. Expert Opin. Drug Discov. 2023, 18, 363–370. [Google Scholar] [CrossRef]
- Won, H.I.; Zinga, S.; Kandror, O.; Akopian, T.; Wolf, I.D.; Schweber, J.T.P.; Schmid, E.W.; Chao, M.C.; Waldor, M.; Rubin, E.J.; et al. Targeted protein degradation in mycobacteria uncovers antibacterial effects and potentiates antibiotic efficacy. Nat. Commun. 2024, 15, 48506. [Google Scholar] [CrossRef]
- Izert, M.A.; Klimecka, M.M.; Górna, M.W. Applications of bacterial degrons and degraders—Toward targeted protein degradation in bacteria. Front. Mol. Biosci. 2021, 8, 669762. [Google Scholar] [CrossRef]
- Ahmad, H.; Zia, B.; Husain, H.; Husain, A. Recent advances in PROTAC-based antiviral strategies. Vaccines 2023, 11, 270. [Google Scholar] [CrossRef]
- Liang, J.; Wu, Y.; Lan, K.; Dong, C.; Wu, S.; Li, S.; Zhou, H.B. Antiviral PROTACs: Opportunity borne with challenge. Cell Insight 2023, 2, 100092. [Google Scholar] [CrossRef] [PubMed]
- Luo, D.; Luo, R.; Wang, W.; Deng, R.; Wang, S.; Ma, X.; Pu, C.; Liu, Y.; Zhang, H.; Yu, S.; et al. Discovery of L15 as a novel Vif PROTAC degrader with antiviral activity against HIV-1. Bioorganic Med. Chem. Lett. 2024, 111, 129880. [Google Scholar] [CrossRef] [PubMed]
- Emert-Sedlak, L.A.; Tice, C.M.; Shi, H.; Alvarado, J.J.; Shu, S.T.; Reitz, A.B.; Smithgall, T.E. PROTAC-mediated degradation of HIV-1 Nef efficiently restores cell-surface CD4 and MHC-I expression and blocks HIV-1 replication. Cell Chem. Biol. 2024, 31, 658–668.e14. [Google Scholar] [CrossRef]
- Mukerjee, N.; Maitra, S.; Mukherjee, D.; Ghosh, A.; Alexiou, A.T.; Thorat, N.D. Harnessing PROTACs to combat H5N1 influenza: A new frontier in viral destruction. J. Med. Virol. 2024, 96, e29926. [Google Scholar] [CrossRef]
- Alugubelli, Y.R.; Xiao, J.; Khatua, K.; Kumar, S.; Sun, L.; Ma, Y.; Ma, X.R.; Vulupala, V.R.; Atla, S.; Blankenship, L.R.; et al. Discovery of first-in-class PROTAC degraders of SARS-CoV-2 main protease. J. Med. Chem. 2024, 67, 6495–6507. [Google Scholar] [CrossRef]
- Desantis, J.; Goracci, L. Proteolysis targeting chimeras in antiviral research. Future Med. Chem. 2022, 14, 459–462. [Google Scholar] [CrossRef]
- Salerno, A.; Seghetti, F.; Caciolla, J.; Uliassi, E.; Testi, E.; Guardigni, M.; Roberti, M.; Milelli, A.; Bolognesi, M.L. Enriching proteolysis targeting chimeras with a second modality: When two are better than one. J. Med. Chem. 2022, 65, 9507–9530. [Google Scholar] [CrossRef]
- Gao, H.; Sun, X.; Rao, Y. PROTAC technology: Opportunities and challenges. ACS Med. Chem. Lett. 2020, 11, 237–240. [Google Scholar] [CrossRef]
- Burslem, G.M.; Smith, B.E.; Lai, A.C.; Jaime-Figueroa, S.; McQuaid, D.C.; Bondeson, D.P.; Toure, M.; Dong, H.; Qian, Y.; Wang, J.; et al. The advantages of targeted protein degradation over inhibition: An RTK case study. Cell Chem. Biol. 2018, 25, 67-77.e3. [Google Scholar] [CrossRef]
- Pu, C.; Wang, S.; Liu, L.; Feng, Z.; Zhang, H.; Gong, Q.; Sun, Y.; Guo, Y.; Li, R. Current strategies for improving limitations of proteolysis targeting chimeras. Chin. Chem. Lett. 2023, 34, 107927. [Google Scholar] [CrossRef]
- Sun, X.; Gao, H.; Yang, Y.; He, M.; Wu, Y.; Song, Y.; Tong, Y.; Rao, Y. PROTACs: Great opportunities for academia and industry. Signal Transduct. Target. Ther. 2019, 4, 64. [Google Scholar] [CrossRef]
- Danishuddin; Jamal, M.S.; Song, K.S.; Lee, K.W.; Kim, J.J.; Park, Y.M. Revolutionizing drug targeting strategies: Integrating artificial intelligence and structure-based methods in PROTAC development. Pharmaceuticals 2023, 16, 1649. [Google Scholar] [CrossRef]
- He, X.; Weng, Z.; Zou, Y. Progress in the controllability technology of PROTAC. Eur. J. Med. Chem. 2024, 265, 116096. [Google Scholar] [CrossRef]
- Rana, S.; Mallareddy, J.R.; Singh, S.; Boghean, L.; Natarajan, A. Inhibitors, PROTACs and molecular glues as diverse therapeutic modalities to target cyclin-dependent kinase. Cancers 2021, 13, 5506. [Google Scholar] [CrossRef] [PubMed]
- Espinoza-Chávez, R.M.; Salerno, A.; Liuzzi, A.; Ilari, A.; Milelli, A.; Uliassi, E.; Bolognesi, M.L. Targeted protein degradation for infectious diseases: From basic biology to drug discovery. ACS Bio Med Chem Au 2023, 3, 32–45. [Google Scholar] [CrossRef]
- Doyle, P.S.; Zhou, Y.M.; Engel, J.C.; McKerrow, J.H. Novel structural CYP51 mutation in Trypanosoma cruzi associated with multidrug resistance to CYP51 inhibitors and reduced infectivity. Int. J. Parasitol. Drugs Drug Resist. 2020, 13, 107–120. [Google Scholar] [CrossRef]
- Lander, N.; Li, Z.H.; Niyogi, S.; Docampo, R. Upregulation of the secretory pathway in cysteine protease inhibitor-resistant Trypanosoma cruzi. J. Cell Sci. 2000, 113, 1345–1354. [Google Scholar] [CrossRef]
- Wurnig, S.; Vogt, M.; Hogenkamp, J.; Dienstbier, N.; Borkhardt, A.; Bhatia, S.; Hansen, F.K. Development of the first geldanamycin-based HSP90 degraders. Front. Chem. 2023, 11, 1219883. [Google Scholar] [CrossRef] [PubMed]
- Pires, D.E.V.; Blundell, T.L.; Ascher, D.B. pkCSM: Predicting small-molecule pharmacokinetic and toxicity properties using graph-based signatures. J. Med. Chem. 2015, 58, 4066–4072. [Google Scholar] [CrossRef]
- González, J.; Ramalho-Pinto, F.J.; Frevert, U.; Ghiso, J.; Tomlinson, S.; Scharfstein, J.; Corey, E.J.; Nussenzweig, V. Proteasome activity is required for the stage-specific transformation of a protozoan parasite. J. Exp. Med. 1996, 184, 1909–1918. [Google Scholar] [CrossRef]
- Bijlmakers, M.J. Ubiquitination and the proteasome as drug targets in trypanosomatid diseases. Front. Chem. 2021, 8, 630888. [Google Scholar] [CrossRef] [PubMed]
- Hashimoto, M.; Murata, E.; Aoki, T. Secretory protein with RING finger domain (SPRING) specific to Trypanosoma cruzi is directed, as a ubiquitin ligase related protein, to the nucleus of host cells. Cell. Microbiol. 2010, 12, 19–30. [Google Scholar] [CrossRef]
- Burge, R.J.; Mottram, J.C.; Wilkinson, A.J. Ubiquitin and ubiquitin-like conjugation systems in trypanosomatids. Curr. Opin. Microbiol. 2022, 70, 102202. [Google Scholar] [CrossRef] [PubMed]
- Cerqueira, P.G.; Passos-Silva, D.G.; Vieira-da-Rocha, J.P.; Mendes, I.C.; de Oliveira, K.A.; Oliveira, C.F.B.; Vilela, L.F.F.; Nagem, R.A.P.; Cardoso, J.; Nardelli, S.C.; et al. Effect of ionizing radiation exposure on Trypanosoma cruzi ubiquitin-proteasome system. Mol. Biochem. Parasitol. 2017, 212, 55–67. [Google Scholar] [CrossRef]
- Ito, T. Protein degraders—From thalidomide to new PROTACs. J. Biochem. 2024, 175, 507–519. [Google Scholar] [CrossRef]
- Yang, X.; Wang, Z.; Pei, Y.; Song, N.; Xu, L.; Feng, B.; Wang, H.; Luo, X.; Hu, X.; Qiu, X.; et al. Discovery of thalidomide-based PROTAC small molecules as the highly efficient SHP2 degraders. Eur. J. Med. Chem. 2021, 218, 113341. [Google Scholar] [CrossRef]
- Bondeson, D.P.; Smith, B.E.; Burslem, G.M.; Buhimschi, A.D.; Hines, J.; Jaime-Figueroa, S.; Wang, J.; Hamman, B.D.; Ishchenko, A.; Crews, C.M. Lessons in PROTAC design from selective degradation with a promiscuous warhead. Cell Chem. Biol. 2017, 25, 78. [Google Scholar] [CrossRef]
- Ulasov, A.V.; Rosenkranz, A.A.; Georgiev, G.P.; Sobolev, A.S. Nrf2/Keap1/ARE signaling: Towards specific regulation. Life Sci. 2021, 291, 120111. [Google Scholar] [CrossRef]
- Suzuki, M.; Otsuki, A.; Keleku-Lukwete, N.; Yamamoto, M. Overview of redox regulation by Keap1–Nrf2 system in toxicology and cancer. Curr. Opin. Toxicol. 2016, 1, 29–36. [Google Scholar] [CrossRef]
- Yao, R.; Luo, T.; Wang, M. Delivering on cell-selective protein degradation using chemically tailored PROTACs. ChemBioChem 2023, 24, e202300413. [Google Scholar] [CrossRef]
- Ishii, M.; Akiyoshi, B. Targeted protein degradation using deGradFP in Trypanosoma brucei. Wellcome Open Res. 2022, 7, 175. [Google Scholar] [CrossRef]
- Poongavanam, V.; Atilaw, Y.; Siegel, S.; Giese, A.; Lehmann, L.; Meibom, D.; Erdelyi, M.; Kihlberg, J. Linker-dependent folding rationalizes PROTAC cell permeability. J. Med. Chem. 2022, 65, 13029–13040. [Google Scholar] [CrossRef]
- Troup, R.I.; Fallan, C.; Baud, M.G.J. Current strategies for the design of PROTAC linkers: A critical review. Explor. Target. Anti-Tumor Ther. 2020, 1, 273–312. [Google Scholar] [CrossRef]
- Poongavanam, V.; Peintner, S.; Abeje, Y.; Kölling, F.; Meibom, D.; Erdelyi, M.; Kihlberg, J. Linker-determined folding and hydrophobic interactions explain a major difference in PROTAC cell permeability. ACS Med. Chem. Lett. 2025, 16, 681–687. [Google Scholar] [CrossRef] [PubMed]
- Li, Y.; Zhang, X.; Xie, J.; Meng, D.; Liu, M.; Chang, Y.; Feng, G.; Jiang, J.; Deng, P. Analyzing the linker structure of PROTACs throughout the induction process: Computational insights. J. Med. Chem. 2025, 68, 3420–3432. [Google Scholar] [CrossRef]
- Zografou-Barredo, N.A.; Hallatt, A.J.; Goujon-Ricci, J.; Cano, C. A beginner’s guide to current synthetic linker strategies towards VHL-recruiting PROTACs. Bioorganic Med. Chem. 2023, 88-89, 117334. [Google Scholar] [CrossRef]
- Abeje, Y.E.; Wieske, L.H.E.; Poongavanam, V.; Maassen, S.; Atilaw, Y.; Cromm, P.; Lehmann, L.; Erdelyi, M.; Meibom, D.; Kihlberg, J. Impact of linker composition on VHL PROTAC cell permeability. J. Med. Chem. 2024, 68, 638–657. [Google Scholar] [CrossRef]
- Yokoo, H.; Naito, M.; Demizu, Y. Investigating the cell permeability of proteolysis-targeting chimeras (PROTACs). Expert Opin. Drug Discov. 2023, 18, 357–361. [Google Scholar] [CrossRef]
- Ge, J.; Hsieh, C.Y.; Fang, M.; Sun, H.; Hou, T. Development of PROTACs using computational approaches. Trends Pharmacol. Sci. 2024, 45, 1162–1174. [Google Scholar] [CrossRef] [PubMed]
- Crummy, E.K.; Caine, E.A.; Mikheil, D.; Corona, C.; Riching, K.M.; Hosfield, C.; Urh, M. Monitoring PROTAC interactions in biochemical assays using Lumit immunoassays. In Methods in Enzymology; Academic Press: Cambridge, MA, USA, 2023; Volume 673, pp. 81–113. [Google Scholar] [CrossRef]
- Li, T.; Zong, Q.; Dong, H.; Ullah, I.; Pan, Z.; Yuan, Y. Non-invasive in vivo monitoring of PROTAC-mediated protein degradation using an environment-sensitive reporter. Nat. Commun. 2025, 16, 57191. [Google Scholar] [CrossRef] [PubMed]
- Liu, X.; Zhang, X.; Lv, D.; Yuan, Y.; Zheng, G.; Zhou, D. Assays and technologies for developing proteolysis targeting chimera degraders. Future Med. Chem. 2020, 12, 1155–1179. [Google Scholar] [CrossRef] [PubMed]
- Carmony, K.C.; Kim, K.B. PROTAC-induced proteolytic targeting. Methods Mol. Biol. 2012, 832, 627–638. [Google Scholar] [CrossRef] [PubMed]
- Gross, P.H.; Sheets, K.J.; Warren, N.A.; Ghosh, S.; Varghese, R.E.; Wass, K.E.; Kadimisetty, K. Accelerating PROTAC drug discovery: Establishing a relationship between ubiquitination and target protein degradation. Biochem. Biophys. Res. Commun. 2022, 628, 68–75. [Google Scholar] [CrossRef]
- Nowak, R.P.; Jones, L.H. Target validation using PROTACs: Applying the four pillars framework. SLAS Discov. 2021, 26, 474–483. [Google Scholar] [CrossRef]
- Rahman, M.; Marzullo, B.; Holman, S.W.; Barrow, M.; Ray, A.D.; O’Connor, P.B. Advancing PROTAC characterization: Structural insights through adducts and multimodal tandem-MS strategies. J. Am. Soc. Mass Spectrom. 2024, 35, 285–299. [Google Scholar] [CrossRef]
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
1 | IC50 = 0.3 μM (T. cruzi epimastigote form) | 4.1955 | [76] | |
2 | IC50 = 0.4 μM (T. cruzi epimastigote form) | 4.5856 | [76] | |
3 | Ki = 2.02 µM IC50 = 1.08 µM (TcFPPS) | −1.6633 | [70] | |
4 | Ki = 1.04 µM IC50 = 0.77 µM (TcFPPS) | −1.2732 | [70] | |
5 | Ki = 0.032µM IC50 = 0.037 µM (TcFPPS) | −0.3744 | [70] | |
6 | DS = −176.92 kcal/mol (TcFPPS) | −2.5697 | [77] | |
7 (R) 8 (S) | DS = −172.72 kcal/mol (TcFPPS) DS = −172.91 kcal/mol (TcFPPS) | −1.5405 | [77] | |
9 | DS = −126.34 kcal/mol (TcFPPS) | −0.4777 | [77] | |
10 | DS = −8.3 kcal/mol (TcFPPS) | — | [78] | |
11 | DS = −8.5 kcal/mol (TcFPPS) | — | [78] |
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
12 | IC50 = 6.9 nM (TcSQS) | 0.8518 | [80] | |
13 | IC50 = 6.4 nM (TcSQS) | 2.8959 | [80] | |
14 | IC50 = 39 nM (TcSQS) | 0.0576 | [81] | |
15 | IC50 = 5 nM (TcSQS) | 0.4477 | [81] | |
16 | IC50 = 21.4 nM (TcSQS) | 0.8378 | [81] | |
17 | IC50 = 11.9 nM (TcSQS) | 1.2279 | [81] | |
18 | IC50 = 22 nM (TcSQS) | 0.3036 | [81] | |
19 | IC50 = 30 nM (TcSQS) | 0.5902 | [81] |
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
38 | IC50 = 89 µM (TcGAPDH) | — | [105] | |
39 | IC50 = 97 µM (TcGAPDH) | — | [105] | |
40 | IC50 = 153 µM (TcGAPDH) | — | [105] | |
41 | Km = 8.24 µM (TcGAPDH) | 2.498 | [106] | |
42 | Km = 11.60 µM (TcGAPDH) | 2.2904 | [106] | |
43 | Km = 11.61 µM (TcGAPDH) | 5.42346 | [106] |
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
83 | 92% inhibition (at 1 mM) (TcTS) IC50 = 0.29 mM (TcTS) | 3.7641 | [168] | |
84 | 82.5% inhibition (at 1 mM) (TcTS) | 2.51262 | [169] | |
85 | 76.5% inhibition (at 1 mM) (TcTS) | 2.0916 | [169] | |
86 | 86.9% inhibition (at 1 mM) (TcTS) DS = −11.1 kcal/mol (TcTS) | 3.96022 | [169] | |
87 | IC50 = 0.23 mM (at 1 mM) (TcTS) Ki = 190 µM (TcTS) | 1.1921 | [169] | |
88 | IC50 = 0.21 Mm (at 1 mM) (TcTS) Ki = 140 µM (TcTS) | 1.1419 | [169] | |
89 | 87.6% inhibition (at 1 mM) (TcTS) DS = −9.6 kcal/mol (TcTS) | 2.0702 | [170] |
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
90 | IC50 = 1.3 nM (cruzain) | 2.8 | [180] | |
91 | IC50 = 0.081 nM (cruzain) | 5.0383 | [180] | |
92 | IC50 = 0.09 nM (cruzain) | 5.7269 | [180] | |
93 | IC50 = 0.2 µM (cruzain) | 4.2163 | [181] | |
94 | IC50 = 0.04 µM (cruzain) | 4.0498 | [182] | |
95 | IC50 = 0.01 µM (cruzain) | 5.1944 | [182] |
Compound | Structure | Affinity Data | LogP | Reference |
---|---|---|---|---|
96 | IC50 = 0.39 µM (T. cruzi) | 1.1981 | [187] | |
97 | Kobs = 1.27 s−1 (at 25 µM) (TcNTR) | 1.9958 | [188] | |
98 | Kobs = 0.82 s−1 (at 25 µM) (TcNTR) | 2.5118 | [188] | |
99 | Kobs = 0.68 s−1 (at 25 µM) (TcNTR) | 2.4028 | [188] | |
100 | Kobs = 0.32 (at 25 µM) Kobs = 0.47 (at 50 µM) (TcNTR) | 0.6472 | [189] |
E3 Ligase Class | Characteristics |
---|---|
RING-domain E3 | Largest class in T. cruzi (67 predicted proteins), mediates direct ubiquitin transfer without a thioester intermediate. |
Cullin-RING Ligases (CRLs) | A subset of RING ligases, forming multiprotein complexes essential for regulated protein degradation. T. cruzi possesses 13 CRLs. |
F-Box E3 Ligases | Associated with the SCF complex (Skp1-Cullin1-F-box), involved in substrate selection for degradation, with eight identified proteins. |
U-Box E3 Ligases | Structurally similar to RING ligases yet lacking metal coordination, these proteins are involved in protein quality control, with five members identified to date. |
HECT-domain E3 | Transfers ubiquitin via a thioester intermediate. Sixteen proteins have been identified, with additional domains such as SPRY and ZnF RBZ. |
Linker Type | Examples | Structural Features | Impact on Permeability | Relevance for T. cruzi |
---|---|---|---|---|
Flexible | Alkyl chains (-CH2-) | High conformational freedom, adaptable. | May increase entropic cost and reduce selectivity. | Facilitates complex formation, although it requires optimization to prevent structural collapse. |
Hydrophilic | PEG (-OCH2CH2-) | Enhances aqueous solubility, increases TPSA. | Often reduces passive diffusion. | Useful for solubility tuning; however, excessive polarity should be avoided. |
Rigid | Triazoles, aromatic linkers | Conformationally constrained, stabilizes ternary complex. | Enhances selectivity; however, it may restrict flexibility. | Ideal for improving protein–protein interactions and metabolic stability. |
Semi-Rigid | Piperazine, piperidine | Balances flexibility and rigidity, reduces PSA. | Promotes intramolecular folding, enhances passive diffusion. | Strong candidate for modulating permeability and metabolic stability. |
Lipophilic | Fused rings, spirocycles | Reduces PSA, increases membrane permeability. | Enhances passive transport; however, it may lead to aggregation. | Can optimize intracellular access if carefully tuned. |
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Rodriguez Gini, A.L.; Souza Tada da Cunha, P.; João, E.E.; Man Chin, C.; dos Santos, J.L.; Serra, E.C.; Benito Scarim, C. TrypPROTACs Unlocking New Therapeutic Strategies for Chagas Disease. Pharmaceuticals 2025, 18, 919. https://doi.org/10.3390/ph18060919
Rodriguez Gini AL, Souza Tada da Cunha P, João EE, Man Chin C, dos Santos JL, Serra EC, Benito Scarim C. TrypPROTACs Unlocking New Therapeutic Strategies for Chagas Disease. Pharmaceuticals. 2025; 18(6):919. https://doi.org/10.3390/ph18060919
Chicago/Turabian StyleRodriguez Gini, Ana Luísa, Pamela Souza Tada da Cunha, Emílio Emílio João, Chung Man Chin, Jean Leandro dos Santos, Esteban Carlos Serra, and Cauê Benito Scarim. 2025. "TrypPROTACs Unlocking New Therapeutic Strategies for Chagas Disease" Pharmaceuticals 18, no. 6: 919. https://doi.org/10.3390/ph18060919
APA StyleRodriguez Gini, A. L., Souza Tada da Cunha, P., João, E. E., Man Chin, C., dos Santos, J. L., Serra, E. C., & Benito Scarim, C. (2025). TrypPROTACs Unlocking New Therapeutic Strategies for Chagas Disease. Pharmaceuticals, 18(6), 919. https://doi.org/10.3390/ph18060919