protein-kinase inhibitors as potential leishmanicidal drugs

185
UNIVERSIDAD COMPLUTENSE DE MADRID FACULTAD DE FARMACIA TESIS DOCTORAL Protein-kinase inhibitors as potential leishmanicidal drugs Inhibidores de proteín-quinasas como potenciales fármacos leishmanicidas MEMORIA PARA OPTAR AL GRADO DE DOCTOR PRESENTADA POR Paula Martínez de Iturrate Sanz Directores Luis Ignacio Rivas López Carmen Belén Gil Ayuso-Gontán Madrid © Paula Martínez de Iturrate Sanz, 2021

Upload: others

Post on 29-Dec-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Protein-kinase inhibitors as potential leishmanicidal drugs

UNIVERSIDAD COMPLUTENSE DE MADRID FACULTAD DE FARMACIA

TESIS DOCTORAL

Protein-kinase inhibitors as potential leishmanicidal drugs

Inhibidores de proteín-quinasas como potenciales fármacos leishmanicidas

MEMORIA PARA OPTAR AL GRADO DE DOCTOR

PRESENTADA POR

Paula Martínez de Iturrate Sanz

Directores

Luis Ignacio Rivas López Carmen Belén Gil Ayuso-Gontán

Madrid

© Paula Martínez de Iturrate Sanz, 2021

Page 2: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 3: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 4: Protein-kinase inhibitors as potential leishmanicidal drugs

UNIVERSIDAD COMPLUTENSE DE MADRID

FACULTY OF PHARMACY/FACULTAD DE FARMACIA

DEPARMENT OF MICROBIOLOGY II/DEPARTAMENTO DE

MICROBIOLOGÍA II

DOCTORAL THESIS/TESIS DOCTORAL

PROTEIN-KINASE INHIBITORS AS POTENTIAL

LEISHMANICIDAL DRUGS

INHIBIDORES DE PROTEÍN-QUINASAS COMO

POTENCIALES FÁRMACOS LEISHMANICIDAS

PhD DISSERTATION BY/ MEMORIA PARA OPTAR AL

GRADO DE DOCTOR PRESENTADA POR

Paula Martínez de Iturrate Sanz

SUPERVISORS/DIRECTORES:

Dr. Luis Ignacio Rivas López

Dr. Carmen Belén Gil Ayuso-Gontán

Madrid, 2020

Page 5: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 6: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 7: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 8: Protein-kinase inhibitors as potential leishmanicidal drugs

Agradecimientos

Estos años de tesis han supuesto una experiencia vital que ha influido mucho en mi

desarrollo personal. Una experiencia que ha formado parte de un puente entre una etapa más

estudiantil e ingenua, y otra más madura y encarada con la vida adulta. Un trabajo que he

podido llevar a cabo gracias a la financiación del Fondo de Garantía Juvenil y a la

financiación del proyecto SAF2015-65740-R (MINECO). En las siguientes líneas, me

gustaría dedicar mis agradecimientos a las distintas personas que me han acompañado en

este viaje.

Me gustaría agradecer en primer lugar a mis directores de tesis, Luis Rivas y Carmen

Gil, por haberme brindado esta oportunidad y haberme dirigido a lo largo del camino.

Muchas gracias por todo lo que me habéis enseñado, y por vuestra paciencia y continua

motivación.

A Mariángeles y Montse quiero agradecerles la figura mentora que fueron para mí

cuando llegué al laboratorio, enseñándome con gran paciencia lo primero que aprendí sobre

Leishmania y técnicas de biología molecular. A Montse quiero agradecerle además su gran

compañerismo, y el enorme apoyo que supuso para mí durante su último año en el

laboratorio. En ella encontré una mentora, compañera y amiga. De la misma manera quiero

darle las gracias a Lorena, por haber sido un pilar clave de apoyo tanto personal como laboral

cuando más lo necesitaba. Tampoco quiero olvidar a aquellas personas que he conocido a su

paso por el laboratorio en estos años, como Sara, Sergio, Javi, Yulieth, Vicky, Iñaqui y

especialmente Cathy, en quien he encontrado una amiga para toda la vida. A todos vosotros

y más, por vuestro compañerismo, por vuestra amistad… muchas gracias.

También quiero agradecerles a aquellas personas que, quizá no lo saben, pero se

convirtieron en un pequeño refugio, una bocanada de aire, especialmente durante mi último

año de trabajo experimental. Por un lado, a Ana María y Concha, así como Gloria, Jose,

Adrián, Ana y Sofi (quien espero que me vuelva a llevar a clases de swing en un futuro tan

cercano como sea posible), con quienes he compartido numerosas conversaciones en la

comida. Y por otro lado, a “los del voley”, por esos buenos ratos de desconexión y deporte,

Page 9: Protein-kinase inhibitors as potential leishmanicidal drugs

y en especial a Guille y Ana, que siempre han tenido palabras de ánimos cuando me han

encontrado en un momento bajo. Gracias.

A Ana Martínez, porque gracias a ella tuve la oportunidad de trabajar con Carmen Gil

y Luis Rivas, y a su grupo por haber generado la biblioteca química que he podido emplear

en este trabajo. Especialmente quiero agradecerle a Víctor Sebastián-Pérez su esencial papel

en la pre-selección virtual de los compuestos y sus estudios computacionales, así como en

la síntesis de los mismos. Muchas gracias.

A mis amigos Nuria y Adrián por haber estado siempre a mi lado, por aguantar esos

audios de Whatsapp interminables y mandarme otros tantos de vuelta, porque nos encanta

irnos por las ramas y sentir que estamos ahí, aun en la distancia. Al “grupo de amigos de

Madrid”, aunque algunos sean de otras ciudades e incluso estén en otros países. Por esos

días de “pizzina”, esos concursos de tapas, esas noches de cine y por más picnics en Debod

(cuando el Sr. Covid lo permita…). Por las risas, las lágrimas, las filosofadas, las tonterías.

Muchas gracias.

A toda mi familia, porque pase lo que pase siempre han estado ahí cuando lo he

necesitado. Gracias a mi prima Ana, porque fue ella quien pensó en mí cuando se enteró de

la convocatoria del CSIC de Garantía Juvenil gracias a la cual tuve mi primer contacto con

el mundo laboral, aquí en el CIB, haciendo lo que más me gusta. Gracias a mi Yaya por

llevarnos siempre a todos en su mente y corazón, sin importar las distancias. Gracias a mis

padres, Mª Pilar y Jesús, por haberme apoyado en todo lo que he necesitado, cada uno a su

propia manera, siempre a mi lado. Gracias a mi hermano Adrián, porque su apoyo fraternal

cala con más fuerza que ningún otro. Gracias. Os quiero infinito.

Finalmente, a mi pareja, Curro. Gracias por estar en lo bueno y en lo malo. Por tu

paciencia y tu comprensión. Por preocuparte. Por animarme. Por apoyarme. Te quiero.

A todos vosotros, una vez más, muchas gracias.

Page 10: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 11: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 12: Protein-kinase inhibitors as potential leishmanicidal drugs

Index

Abbreviations ....................................................................................................................... 1

Resumen ............................................................................................................................... 7

Summary ............................................................................................................................ 11

1. Introduction ................................................................................................................... 17

1.1. Leishmaniasis ............................................................................................................. 17

1.1.1. Taxonomy of Leishmania ....................................................................................... 18

1.1.2. Clinical manifestations of leishmaniasis .................................................................. 20

1.1.3. Life cycle of Leishmania ........................................................................................ 21

1.1.4. Control of leishmaniasis ......................................................................................... 23

1.1.4.1. Prevention of leishmaniasis infection: vaccines, vector and reservoir control ..... 23

1.1.4.2. Diagnosis ....................................................................................................... 24

1.1.5. Biology of Leishmania ........................................................................................... 25

1.1.5.1. Metabolic characteristics of Leishmania .......................................................... 25

1.1.5.1.1. Carbon metabolism ................................................................................... 25

1.1.5.1.2. Oxidative phosphorylation ......................................................................... 27

1.1.5.1.3. Other metabolic peculiarities: Folate synthesis, ergosterol and trypanothione.... 29

1.1.6. Signal transduction in Leishmania .......................................................................... 29

1.1.6.1. The kinome of Leishmania ............................................................................... 31

1.1.7. State of the art of the chemotherapy for leishmaniasis .............................................. 39

2. Objectives ...................................................................................................................... 47

3. Materials and Methods ................................................................................................. 51

3.1. Chemical compounds, media, reagents and laboratory equipment ............................ 51

3.2. Cell cultures and cell harvesting ................................................................................ 51

3.3. Leishmanicidal activity of the compounds on axenic parasites ................................. 52

Page 13: Protein-kinase inhibitors as potential leishmanicidal drugs

3.4. Macrophage cytotoxicity of leishmanicidal compounds ........................................... 53

3.5. Leishmanicidal activity against intracellular amastigotes ......................................... 54

3.6. Extraction and purification of LdGSK-3s .................................................................. 54

3.7. Measurement of LdGSK-3s activity .......................................................................... 56

3.8. Bioenergetic assays .................................................................................................... 57

3.8.1. In vivo monitoring of intracellular ATP levels ......................................................... 57

3.8.2. Assessment of plasma membrane permeabilization .................................................. 57

3.8.3. Mitochondrial membrane potential (ΔΨm) of Leishmania ......................................... 58

3.8.4. Evaluation of the O2 consumption rate of the parasite .............................................. 58

3.8.5. Assessment of programed cell death induced by the compounds ............................... 59

4. Results and discussion .................................................................................................. 63

4.1. LdGSK-3s extraction and purification ....................................................................... 65

4.2. Selection of the different compounds tested .............................................................. 70

4.3. Assessment of the biological and inhibitory activities of the selected compounds ... 71

4.3.1. Reported inhibitors of human protein kinases (Sets 1 and 1.1) .................................. 72

4.3.2. In silico inhibitors of LmjGSK-3s (Set 2), LmjCK1.2 (Set 3) and LmxMAPK4 (Set 4) ..

............................................................................................................................ 89

4.3.3. Leishmanicidal activity of LdGSK-3s inhibitors on intracellular amastigotes ........... 114

4.4. Off-target effects of protein kinase inhibitors .......................................................... 117

4.4.1. Energy metabolism of Leishmania as an off-target effect of PKIs ........................... 118

4.4.1.1. Variation of the intracellular levels of ATP in L. donovani promastigotes by the

leishmanicidal LdGSK-3s inhibitors .......................................................................... 119

4.4.1.2. Inhibition of the electrochemical potential of the Leishmania mitochondrion (ΔΨm)

................................................................................................................... 122

4.4.1.3. Inhibition of O2 consumption ......................................................................... 124

4.4.1.4. Identification of the PKI target inside the respiratory chain ............................ 127

4.4.1.5. Induction of programmed cell death in L. infantum promastigotes ................... 134

Page 14: Protein-kinase inhibitors as potential leishmanicidal drugs

5. Conclusions .................................................................................................................. 139

Bibliography ..................................................................................................................... 143

Results dissemination ...................................................................................................... 167

Page 15: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 16: Protein-kinase inhibitors as potential leishmanicidal drugs

Abbreviations

1

Abbreviations

α-GP α-glycerophosphate

αKG α-ketoglutarate

ΔΨm mitochondrial electrochemical potential

λEM excitation wavelength

λEX emission wavelength

5-Me-6-BIO 6-bromo-5-methylindirubin-3’-oxime

6-BIO 6-bromoindirubin-3’-oxime

Ac acetate

AC adenylate cyclase

AcCoA acetyl-CoA

ADP adenosine diphosphate

AGC kinase group that includes the protein kinase A, G and C

families

AIDS acquired immune deficiency syndrome

AIRK aurora kinase

ALM autoclaved-killed L. major vaccine prototype

AmB amphotericin B

AmBd amphotericin B deoxycholate

Asc ascorbate

ATP adenosine triphosphate

CAMK Ca2+/calmodulin-dependent kinase

CAMK1 Ca2+/calmodulin-dependent kinase 1

CAMKL CAMK-like kinase

cAMP cyclic adenosine monophosphate

cGMP cyclic guanosine monophosphate

CDK/CRK cyclin-dependent kinase/cdc2-related kinase

CDPK Ca2+-dependent protein kinase

Cit citrate

CK1 casein kinase 1

CK2 casein kinase 2

Page 17: Protein-kinase inhibitors as potential leishmanicidal drugs

2

CL cutaneous leishmaniasis

CLK cell division control (CDC)-like kinase

CMGC kinase group that includes the CDK, MAPK, GSK and CLK

families

CoQ Coenzyme Q

CytC cytochrome c

DALY disability-adjusted life years

Dig digitonin

DMNPE D-luciferin D-Luciferin 1-(4,5-dimethoxy-2-nitrophenyl)ethyl ester

DNDi Drugs for Neglected Diseases initiative

DYRK dual-specificity tyrosine-regulated kinase

EDTA ethylenediaminetetraacetic acid

ePK eukaryotic protein kinase

FAD flavin adenine dinucleotide

FCCP carbonyl cyanide-p-trifluoromethoxyphenylhydrazone

FDA Food and Drug Administration

FFAA fatty acids

FRD fumarate reductase

G3P glyceraldehyde 3-phosphate

Glc6P glucose 6-phosphate

GALM gentamycin-attenuated L. major vaccine prototype

GPCR G-protein-coupled receptor

GS2 phosphoglycogen synthase peptide-2

GSK-3 glycogen-synthase kinase-3

HBSS Hank’s balanced salt solution

HePc hexadecylphosphocholine (miltefosine)

HEPES 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid

HIFCS heat inactivated fetal calf serum

HIV Human Immunodeficiency Virus

HMK halomethylketone

hPK human protein-kinase

HPLC high-performance liquid chromatography

Page 18: Protein-kinase inhibitors as potential leishmanicidal drugs

Abbreviations

3

IC50 compound concentration that induces 50% inhibition

IC80 compound concentration that induces 80% inhibition

i.m. intramuscular

i.v. intravenously

IPTG isopropyl β-D-1-thiogalactopyranoside

ITDZ iminothiadiazole

kDNA kinetoplast

LAmB liposomal amphotericin B

LAMP loop-mediated isothermal amplification

LRRK2 leucine-rich repeat kinase 2

M199 medium 199

M199-HIFCS M199 supplemented with 20% HIFCS

MA meglumine antimoniate

Mal malate

MAPK mitogen-activated protein kinase

MAPKK mitogen-activated protein kinase kinase

MAPKKK mitogen-activated protein kinase kinase kinase

MAPKKKK mitogen-activated protein kinase kinase kinase kinase

MBC medicinal and biological chemistry

MCL mucocutaneous leishmaniasis

MPM murine peritoneal macrophages

MTT 3-[4,5-dimethylthiazole-2-yl]-2,5-diphenyltetrazolium

bromide

MW molecular weight

NAD nicotinamide adenine dinucleotide

NDR nuclear DBF2-related kinase

NEK never in mitosis gene A(NimA)-related kinase

OAA oxaloacetate

OD optical density

Omy oligomycin

PBS phosphate-buffered saline

PCR polymerase chain reaction

Page 19: Protein-kinase inhibitors as potential leishmanicidal drugs

4

PDK1 phosphoinositide-dependent protein kinase 1

PEP phosphoenolpyruvate

PI propidium iodide

PK protein kinase

PKDL post-kala azar dermal leishmaniasis

PKA protein kinase A; cAMP-dependent protein kinase

PKB protein kinase B; Akt kinase

PKC protein kinase C

PKG protein kinase G; cGMP-dependent protein kinase

PKI protein kinase inhibitor

PLK polo-like kinase

PPP pentose phosphate pathway

psi pounds per square inch

Pyr pyruvate

RCK ROS-cross-hybridizing kinase

RGC Receptor guanylate cyclase group

RNS reactive nitrogen species

ROS reactive oxygen species

RPMI 1640 Roswell Park Memorial Insitute 1640 medium

RPMI-HIFCS RPMI 1640 medium supplemented with 10% HIFCS

RPMIØRED RPMI 1640 medium without red phenol

RPMIØRED-HIFCS RPMIØRED medium supplemented with 10% HIFCS

RSK ribosomal S6-kinase

SAR structure-activity relationship

SCoA succinyl-CoA

SDS-PAGE sodium dodecyl sulfate-polyacrylamide gel electrophoresis

SI selectivity index

SRPK SR-rich protein kinase

SSG sodium stibogluconate

STE kinase homolog of yeast sterile

STE7 kinase homolog of yeast sterile 7

STE11 kinase homolog of yeast sterile 11

Page 20: Protein-kinase inhibitors as potential leishmanicidal drugs

Abbreviations

5

STE20 kinase homolog of yeast sterile 20

Succ succinate

TCA tricarboxylic acid cycle

TDZD thiadiazolidindione

TK tyrosine kinase

TKL tyrosine kinase-like

TMPD N,N,N',N'-tetramethyl-p-phenylenediamine

TPP target product profile

TTBK tau tubulin kinase

TX-100 Triton X-100

VL visceral leishmaniasis

WEE1 Wee1-like protein kinase 1

WHO World Health Organization

Page 21: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 22: Protein-kinase inhibitors as potential leishmanicidal drugs

Resumen

7

Resumen

Título: Inhibidores de proteín-quinasas como potenciales fármacos leishmanicidas.

Introducción

La leishmaniasis es una enfermedad parasitaria de transmisión vectorial causada por

hasta 20 especies del género Leishmania, un protozoo con un ciclo de vida caracterizado por

tener dos formas morfológicas adaptadas a cada uno de sus dos hospedadores: el

promastigote (extracelular) en el vector invertebrado, y el amastigote (intracelular) en el

hospedador vertebrado. Aunque la transmisión de la leishmaniasis es principalmente

zoonótica, el ciclo antroponótico de la enfermedad tiene lugar en zonas endémicas con alta

densidad humana. Su prevalencia ha aumentado en las últimas décadas debido a diversos

factores incluyendo el aumento de viajes por todo el mundo y el aumento de la distribución

geográfica del vector flebotomo por el calentamiento global.

Aunque hay vacunas humanas en desarrollo, el tratamiento principal es la

quimioterapia. Sin embargo, los fármacos en uso presentan problemas de costes, toxicidad

y aparición de resistencias. En la búsqueda de nuevos tratamientos farmacológicos se están

llevando a cabo dos estrategias: desarrollo de tratamientos combinados con los actuales

fármacos disponibles, y descubrimiento de nuevos compuestos con potencial leishmanicida

mediante tres posibles acercamientos: reposicionamiento de fármacos, cribados fenotípicos

y cribados dirigidos a diana.

En Leishmania hay pocas dianas validadas como esenciales para el parásito, entre ellas

las proteín-quinasas GSK-3s, CK1.2 y MAPK4. Las proteín-quinasas conforman una de las

familias de proteínas más numerosas en la mayoría de organismos eucariotas, incluida

Leishmania, dada su participación en numerosos procesos celulares, muchos de ellos

esenciales. Por ello, los cribados enfocados en una proteín-quinasa validada son de gran

interés, especialmente si se combinan con un cribado fenotípico que aseguren su efectividad

en el parásito, y permitan la búsqueda de actividades leishmanicidas más allá de la diana

específica objeto del cribado.

Page 23: Protein-kinase inhibitors as potential leishmanicidal drugs

8

Objetivos generales

El principal objetivo de esta tesis se ha centrado en el descubrimiento de nuevos

agentes leishmanicidas con un mecanismo de acción definido. Para ello, se han evaluado

distintos grupos de compuestos orgánicos de bajo peso molecular mediante la combinación

de un cribado enfocado en la inhibición de la isoforma corta de la GSK-3 de L. donovani

(LdGSK-3s) y un cribado fenotípico en promastigotes y amastigotes axénicos. Además, se

evaluó el efecto de los compuestos activos en distintos parámetros bioenergéticos para

definir su efecto en la homeostasis intracelular de parásito y explorar posibles dianas

adicionales.

Resultados

Tras expresar y purificar la GSK-3s de L. donovani a partir de E. coli BL21(D3)

transfectadas con un plásmido pET28a(+) con el gen de dicha proteín-quinasa, se

seleccionaron 4 grupos de compuestos (Sets 1 a 4). El Set 1 se compuso con inhibidores de

proteín-quinasas humanas, mientras que los Sets 2, 3 y 4 se originaron a partir de hits

obtenidos de cribados virtuales enfocados en la GSK-3s y la CK1.2 de L. major y la MAPK4

de L. mexicana, respectivamente. Adicionalmente, se formaron los Sets 1.1 y 2.1 con

compuestos seleccionados a partir de la optimización de los mejores compuestos activos de

los Sets 1 a 4.

En total fueron evaluados 136 compuestos, y se identificaron 14 hits con actividad

inhibidora en LdGSK-3s y actividad leishmanicida, pertenecientes a tres familias químicas

diferentes (tiadiazolidindionas, halometilcetonas y quinonas). De estos 14 compuestos, 5

mostraron un buen índice de selectividad (IS >10). Adicionalmente, 4 quinonas identificadas

(compuestos 72, 75, 76 y 80) son los primeros inhibidores específicos de la GSK-3s de

Leishmania, sin inhibición cruzada con la GSK-3β humana, su homólogo humano más

cercano.

Mediante el uso de herramientas computacionales, se identificó el modo de unión de

las tiadiazolidindionas (compuestos 1 y 2) y las halometilcetonas (compuestos 6, 28, 29, 32

y 33) activas en LdGSK-3s a través de unión covalente a la Cys169. Por otro lado, se evaluó

Page 24: Protein-kinase inhibitors as potential leishmanicidal drugs

Resumen

9

la efectividad de dichos compuestos en macrófagos infectados con amastigotes, como

modelo más cercano al fisiológico. De los 14 hits, las halometilcetonas 6 y 32 y las quinonas

72, 76 y 79 disminuyeron significativamente la carga parasitaria de los macrófagos.

Finalmente, se analizó el efecto de los inhibidores de LdGSK-3s leishmanicidas en el

metabolismo energético de promastigotes de Leishmania. La halometilcetona 6 y las

quinonas 72, 75 y 76 produjeron colapso bioenergético en el parásito, con despolarización

de la membrana mitocondrial, inhibición de la respiración con disminución de los niveles

intracelulares de ATP, y desencadenando finalmente la apoptosis del parásito.

Específicamente, se identificó al complejo III de la cadena transportadora de electrones

como la diana de inhibición del compuesto 6, mientras que las quinonas 72, 75 y 76

mostraron un doble efecto desacoplador e inhibidor de la cadena respiratoria de Leishmania.

Conclusiones

La combinación de un cribado enfocado en la GSK-3s de Leishmania con un cribado

fenotípico en el parásito ha llevado a la identificación de nuevos compuestos de interés para

el desarrollo de nuevos fármacos leishmanicidas. Además, para algunos de estos compuestos

se ha identificado una diana adicional en la cadena respiratoria del parásito. Esta naturaleza

multi-diana es una característica ventajosa que reduce la probabilidad de que emerjan

resistencias causadas por mutaciones en la diana inicial.

Page 25: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 26: Protein-kinase inhibitors as potential leishmanicidal drugs

Summary

11

Summary

Title: Protein-kinase inhibitors as potential leishmanicidal drugs.

Introduction

Leishmaniasis is a parasitic vector-borne disease caused by more than 20 species from

the genus Leishmania, a protozoan with two distinct morphological forms throughout its life

cycle adapted to each one of its two hosts: the promastigote (extracellular) in the invertebrate

vector, and the amastigote (intracellular) in the vertebrate host. Although it is mostly a

zoonotic disease, anthroponotic transmission occurs in endemic areas with high human

density. Moreover, its prevalence increased in the last years due to several factors including

an increase in worldwide travelling as well as an increase in the geographical distribution of

the phlebotomine vector due to global warming.

Despite the efforts to develop human vaccines, the main treatment against

leishmaniasis relies on chemotherapy. However, all the current clinical drugs available have

drawbacks due to toxicity, high-cost and/or resistance. Two approaches are leading the

search for new pharmacological treatments: combination therapy using current drugs

available, and searching for new compounds with a leishmanicidal potential through three

major strategies: drug repurposing, phenotypic screenings and target-based screenings.

Protein kinases GSK-3s, CK1.2 and MAPK4 are among the scarce validated targets in

Leishmania. Owing to their participation in many and diverse cellular processes, protein

kinases are one of the largest protein families in most eukaryotic organisms, including

Leishmania. Hence, target-based screenings focused on a validated protein kinase are of

great interest, especially if combined with a phenotypic screening to ensure effectiveness on

living parasites, as well as to uncover off-target effects.

Page 27: Protein-kinase inhibitors as potential leishmanicidal drugs

12

Objectives

The main objective of this thesis was to find new leishmanicidal agents with a defined

mechanism of action. For this purpose, different groups of small-size organic compounds

have been evaluated by combination of a target-based approach on the inhibition of the short

isoform of the GSK-3 of L. donovani (LdGSK-3s), together with phenotypic screenings on

axenic promastigotes and amastigotes. Furthermore, the effect of active compounds on

different bioenergetic parameters was assessed as a subtle approach to gauge their effect on

the intracellular homeostasis of the parasite and explore feasible off-targets.

Results

The GSK-3s of L. donovani was successfully expressed and purified from E. coli

BL21(D3) transfected with a pET28a(+) plasmid encoding the gene for this protein kinase.

Then, 4 groups of compounds were gathered (Sets 1 to 4). Set 1 comprised human protein

kinase inhibitors, whereas Sets 2, 3 and 4 originated from virtual screenings focused on the

GSK-3s and CK1.2 of L. major, and the MAPK4 of L. mexicana, respectively. Additionally,

Sets 1.1 and 2.1 comprehended compounds selected as a refinement of the best active

compounds from Sets 1 to 4.

Among the 136 compounds evaluated, 14 leishmanicidal hits with inhibitory activity

on LdGSK-3s were identified. They belonged to three different chemical families with

different scaffolds (thiadiazolidindiones, halomethylketones and quinones). Out of these 14

hits, 5 showed a good selectivity index (SI >10). In addition, 4 quinones (compounds 72, 75,

76 and 80) are the first GSK-3 inhibitors of Leishmania, devoid of cross-inhibition on the

human GSK-3β, the closest human homologue for this kinase.

Through computational tools, the interactions of the thiadiazolidindiones (compounds

1 and 2) and halomethylketones (compounds 6, 28, 29, 32 and 33) with LdGSK-3s were

analysed. Formation of a covalent bond with Cys169 was identified as the binding mode of

these hits. Additionally, all 14 hits were evaluated on infected macrophages in order to assess

their effectivity against Leishmania in an intracellular model. Halomethylketones 6 and 32,

and quinones 72, 76 and 79 significantly reduced the parasite load of infected macrophages.

Page 28: Protein-kinase inhibitors as potential leishmanicidal drugs

Summary

13

Finally, the effect of these leishmanicidal LdGSK-3s inhibitors on the energy

metabolism of Leishmania promastigotes was assessed. Halomethylketone 6 and quinones

72, 75 and 76 induced the bioenergetic collapse of the parasite, with mitochondrial

depolarization, inhibition of respiration and decrease of intracellular ATP levels, with

apoptosis ensuing. Their specific targets on the respiratory chain of Leishmania were

complex III for compound 6, while quinones 72, 75, and 76 showed both uncoupling and

inhibitory effects on the respiration of the parasite.

Conclusions

The combination of a target-based screening focused on the GSK-3s of Leishmania

with a phenotypic screening in the parasite has led to the identification of new compounds

with appealing activities for their further development as new leishmanicidal drugs.

Moreover, some of these compounds have an additional target in the respiratory chain of the

parasite. The multi-target nature of these compounds is an advantageous trait to confront a

feasible raise of resistance caused by mutation in a single target.

Page 29: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 30: Protein-kinase inhibitors as potential leishmanicidal drugs

INTRODUCTION

Page 31: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 32: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

17

1. Introduction

1.1. Leishmaniasis

Leishmaniasis is a term that encompasses the group of vector-borne parasitic diseases

caused by more than 20 protozoan species belonging to the genus Leishmania. The parasite

is transmitted by phlebotomine sandflies, and leads to one of three major groups of

leishmaniasis with diverse clinical manifestations: cutaneous (CL), mucocutaneous (MCL),

and visceral leishmaniasis (VL) or kala-azar. It is mostly a zoonotic disease where feral and

domestic mammals act as the main reservoirs of the parasite, but anthroponotic transmission

of Leishmania can take place in areas of high human density.1,2 The World Health

Organization’s (WHO) latest report3 states that, as of 2018, leishmaniasis is endemic in 98

countries across Latin America, the Indian Subcontinent, Central and South America, as well

as South Sudan, Ethiopia, Kenya and European countries of the Mediterranean Region, with

sporadic outbreaks every 10-15 years due to the loss of herd immunity, which is especially

relevant in anthroponotic areas.4 Two of the latest and most relevant sporadic outbreaks in

Southern Europe took place in Fuenlabrada (Spain, 2012)5,6 and Modena (Italy, 1997-2016),7

with more than 400 and 35 immunocompetent patients affected respectively.

Like other poverty-associated diseases, leishmaniasis is one of the so-called Neglected

Tropical Diseases due to its impact on human health and poor economical investment for its

control and treatment. In fact, the highest prevalence of this disease occurs in low-income

populations from developing countries, where people suffer from hunger, poor living

conditions and faulty health facilities. On a global scale, over 1 billion people are at risk each

year, with an estimation of 700,000-1,000,000 new cases of leishmaniasis per year and

26,000-65,000 deaths annually.8 Furthermore, the economic impact of the disease is

estimated in approximately 1 million disability-adjusted life years (DALYs), this is,

1 million years of productive life lost due to disability.9 This number increases up to ca

Page 33: Protein-kinase inhibitors as potential leishmanicidal drugs

18

2.2 million additional DALYs when major depressive disorders associated are included,

accounting for the psychosocial impact of facial mutilation or scarring on CL and MCL

patients.10 Moreover, the impact of leishmaniasis is worsened by co-morbidity with other

pathologies, including malnutrition and infections that compromise the immune system.11,12

Leishmania-HIV coinfection is especially relevant, as this virus targets macrophages and

CD4+ lymphocytes ―the axis of immunity against Leishmania13― with synergetic effects

on the outcome of both infections.14,15

In the last years, the prevalence of leishmaniasis has increased due to several factors

such as “imported leishmaniasis”, associated with the increase in worldwide travelling

(including animal reservoirs and undetected transport of vectors),16-18 as well as human

migrations such as those forced by war conflicts.19,20 It is also inevitably linked to global

warming, which increases the geographical distribution of sandfly species towards higher

latitudes,21,22 as well as loss of efficacy of current treatments due to resistance, side-effects

and/or high costs.23-25

There are several potential human vaccines against leishmaniasis under different

phases of clinical trials, but none are commercially available so far.26 Therefore, the fight

against the spread and persistence of Leishmania is, nowadays, almost exclusively

dependent on chemotherapy despite treatment failures and costs. Aiming to improve this

scenario, the WHO, big pharma companies, and other institutions including academia, are

currently collaborating to make their large collections of compounds available to the

scientific community, as well as to supply current drugs at cost for underdeveloped

countries.27,28

1.1.1. Taxonomy of Leishmania

The taxonomy of Leishmania has undergone multiple reviews,29 from pure

morphological and clinical criteria to genomic comparisons. Currently, the genus

Leishmania belongs to the Trypanosomatidae family, and is divided into three subgenera:

L. (Leishmania), L. (Viannia) and L. (Sauroleishmania). The latter only infects reptiles,

Page 34: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

19

whereas different species within L. (Leishmania) and L. (Viannia) are grouped in complexes

associated with specific clinical forms of leishmaniasis30 (Figure 1).

Figure 1. Taxonomical classification of Leishmania (Modified from van der Auwera et al.31).

Page 35: Protein-kinase inhibitors as potential leishmanicidal drugs

20

1.1.2. Clinical manifestations of leishmaniasis

The diversity of symptoms for leishmaniasis is mostly dependent on the

immunological status of the host and the infecting Leishmania species. From a clinical

perspective, the different pathologies are grouped under three major forms:8

1) Visceral leishmaniasis is the deadliest form of leishmaniasis if left untreated.32 It is

caused by L. donovani and L. infantum in the Old World, and L. chagasi (infantum) in

the New World. Incubation period ranges from 10 days to over 1 year, and symptoms

and clinical signs usually appear gradually. VL is characterised by irregular fever

episodes, weight loss, hepatosplenomegaly and anaemia. For a certain percentage of

immunocompetent VL patients, infection may course as asymptomatic.33 There are

two forms of VL:

Anthroponotic VL or kala-azar is caused by L. donovani in India, Nepal,

Bangladesh, Sudan and Ethiopia.32 Occasionally, 6-12 months or more after an

apparent recovery from VL, patients may suffer a cutaneous complication

denominated post-kala azar dermal leishmaniasis (PKDL). PKDL is

characterised by hypo-pigmented macular, maculopapular, and nodular rash.34 In

Sudan, this relapse may occur earlier or even concurrently to VL, and while PKDL

frequently heals spontaneously in Africa, this rarely happens in India.

Zoonotic VL is caused by L. infantum in Central and South America, and the

Mediterranean Region.8

2) Cutaneous leishmaniasis is the most common form of leishmaniasis. It is caused by

L. aethiopica, L. major, and L. tropica in the Old World, and L. amazonensis and L.

mexicana in the New World. It is characterised by a macule at the site of inoculation

that evolves into a papule.35,36 In a later stage, the papule frequently turns into an ulcer

that leaves a permanent scar after healing. Sometimes, standard CL results in more

complicated forms:37

Diffuse CL derives from CL caused by L. aethiopica in the Old World, and from

L. mexicana and L. amazonensis in the New World. It is characterised by widely

Page 36: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

21

dispersed skin lesions, usually without ulceration. Diffuse CL does not heal

spontaneously and relapses are usual.

Disseminated CL originates in the New World from CL caused by L. braziliensis,

L. panamensis, L. guyanensis and L. amazonensis. Disseminated CL cases show 20

to hundreds of nodular or ulcerated lesions, with or without mucosal damage.

3) Mucocutaneous leishmaniasis occurs exclusively in the New World. It is caused by

species from the braziliensis complex, especially L. braziliensis and L. panamensis.

Leishmania migrates from a primary cutaneous lesion to the mucosal tissues of the

mouth and upper respiratory tract via lymphatic or haematogenous dissemination,

where it causes the complete or partial destruction of the oropharyngeal mucosa and

cartilage.36,38

As stated before, the concurrence of leishmaniasis and AIDS is of special relevance.

Leishmania-HIV co-infected individuals have a higher chance to suffer acute and fatal forms

of leishmaniasis because of their compromised immune system. VL-HIV patients show

atypical organ-infections such as the gastrointestinal tract, peritoneal space, lungs, pleural

space, and skin. In the New World, CL-HIV patients suffer numerous polymorphic and

relapsing lesions.37

1.1.3. Life cycle of Leishmania

Leishmania is a unicellular parasite with a digenetic life cycle and two major

morphological forms, adapted for two extremely different environments.39 The elongated

and flagellated extracellular promastigote faces slightly alkaline conditions and

environmental temperature inside the gut of the sandfly vector. In contrast, the ovoid

non-motile intracellular amastigote dwells in an acidic environment inside the

parasitophorous vacuole of its vertebrate host, full of different hydrolases and reactive

species of oxygen and nitrogen (ROS and RNS respectively), with temperatures ranging

from 32 ºC (CL and MCL) to 37 ºC (VL)40,41 (Figure 2).

Page 37: Protein-kinase inhibitors as potential leishmanicidal drugs

22

Figure 2. Life Cycle of Leishmania (adapted from Hurrel et al.).42

Leishmania vectors are female sandflies from the genus Phlebotomus in the Old World

and Lutzomyia in the New World. Phlebotomine sandflies feed on plant juices and sugary

secretions, but females also blood feed on vertebrate hosts to produce eggs.43 The infected

female sandflies inject metacyclic Leishmania promastigotes from her proboscis into the bite

site during the blood meal44 (1). In the vertebrate host, these metacyclic promastigotes are

phagocytised by neutrophils. Leishmania survives in the neutrophil preventing the fusion of

the phagosome with the azurophil granules, while inducing apoptosis in these cells (2). The

apoptotic neutrophil ensures the uptake of the parasite by the macrophage without promoting

its activation. Alternatively, the metacyclic promastigotes enter the macrophage by

opsonization through the binding of their surface proteins with the C3 factor of the

complement (3). Promastigotes end up in a phagosome that eventually fuses with lysosomes

and finally becomes a parasitophorous vacuole.45 Afterwards, the promastigotes transform

into amastigotes, with an oval shape and a short and afunctional flagellum (4). Amastigotes

Page 38: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

23

multiply within the parasitophorous vacuole until they lyse the macrophage, and continue to

infect and replicate in new macrophages46 (5). When a sandfly feeds on a Leishmania-

infected vertebrate, they ingest infected macrophages and/or circulating amastigotes (6).

Amastigotes revert to promastigotes in the gut of the vector and, after several rounds of

replication, undergo metacyclogenesis (7). Through metacyclogenesis, initial low-infective

procyclic promastigotes migrate into the proboscis as highly infective metacyclic

promastigotes, closing the cycle.47

1.1.4. Control of leishmaniasis

The epidemiology of leishmaniasis is determined by many different factors that

influence vector and human behaviours, such as climate change and war conflicts.48,49

Specific protocols in each regional context have been established to address the persistence

and prevalence of leishmaniasis. While early diagnosis and chemotherapy are essential to

alleviate the severity of the pathology and prevent death, vaccines as well as vector and

reservoir control are useful to prevent infection and reduce the transmission of Leishmania.50

1.1.4.1. Prevention of leishmaniasis infection: vaccines, vector and reservoir control

Although it is well known that immunoprotection is achieved after recovery from

leishmaniasis, a human vaccine has not been approved yet. Three generations of vaccines

have been developed over the past few decades. First generation vaccines consist of whole

parasites physically/genetically attenuated or killed. Second generation vaccines are

developed with specific protein antigens and the third generation includes genetic

immunization through an expression vector encoding parasite antigens to be expressed by

the host.50-52 So far, few vaccines have entered human clinical trials, and none have reached

Phase III yet (Table 1). The latest vaccine to enter clinical trials was the third-generation

vaccine ChAd63-KH, which was designed for prevention and treatment of VL and PKDL

caused by L. donovani,26 and consists of simian adenovirus ChAd63 carrying two

Leishmania genes that encode two parasite proteins (KMP11 and HASPB).50

Page 39: Protein-kinase inhibitors as potential leishmanicidal drugs

24

Table 1. State-of-art of human vaccines against leishmaniasis under clinical trials.26

Clinical Phase I Clinical Phase II

First generation

vaccine -

Autoclaved-killed L. major (ALM)

Gentamycin-attenuated L. major (GALM)

Second generation

vaccine

LEISH-F1

LEISH-F3

LEISH-F2

SMT+NH

Third generation

vaccine - ChAd63-KH

Despite the current efforts to develop human vaccines, nowadays prevention of

leishmaniasis essentially relies upon the control of reservoirs and vectors.48,53 Sandfly

control encompasses physical and chemical barriers: use of indoor and outdoor insecticide

sprays, insecticide-impregnated curtains and bed nets, topical insect repellents, and

minimization of skin exposure. Options for reservoir control ranges from dog collar

repellents, canine drugs and vaccines for domestic dogs, to destruction or chemical treatment

of burrows of wild animals, as well as euthanasia and culling.54-56

1.1.4.2. Diagnosis

Early diagnosis is key to mitigating the development of a more severe pathology and

even death. Current standardized protocols include clinical, parasitological, serological and

antigen-detection tests depending on the type of leishmaniasis and resources available on the

spot.57,58 Most tests require special equipment only available in hospitals and research

centres, but there are two diagnostic tests available on the field: the direct agglutination test,

and the rK39 dip-stick test. The immunochromatographic rK39 test has proven to be

essential for early diagnosis of VL despite its poor performance in VL-HIV and PKDL

patients and its inability to discriminate asymptomatic infections.52 New methods in the

pipeline for early detection of leishmaniasis include the loop-mediated isothermal

amplification (LAMP) assay, a PCR assay that works at environmental temperature.59

Page 40: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

25

1.1.5. Biology of Leishmania

In comparison to other eukaryotic cells, trypanosomatids show some peculiarities

concerning their intracellular organelles, such as glycosomes, acidocalcisomes, and a single

well-developed mitochondrion, accounting for significant metabolic differences:60

― Glycosomes are peroxisome-like organelles found in trypanosomatids that harbour the

enzymes involved in the first seven steps of glycolysis, as well as enzymes for other

metabolic pathways such as the pentose phosphate pathway (PPP), gluconeogenesis and

part of the β-oxidation of fatty acids.61

― Acidocalcisomes are membrane-bound metal-storage organelles implicated in

polyphosphate and pyrophosphate metabolism, also found in Apicomplexa parasites.62

― All Kinetoplastida including Trypanosomatidae possess a single mitochondrion with

extensive ramification that accounts for up to 12% of the volume of the parasite. In

contrast, mammalian cells can encompass hundreds of mitochondria in as much as 20%

of their cell volume. Furthermore, the trypanosomatid mitochondrion contains a unique

well-developed mitochondrial disk-shaped DNA called kinetoplast (kDNA) that

accounts for about 30% of the total DNA of these parasites.63,64

1.1.5.1. Metabolic characteristics of Leishmania

In trypanosomatid parasites, adaptive mechanisms to survive in their hosts include

energy and redox balance, regulated by the compartmentalization of different metabolic

pathways within the cytoplasm and the aforementioned organelles.65 Moreover, Leishmania

has stage-specific metabolic adaptations aimed to ensure an adequate supply of nutrients and

evade different stresses in each of its two hosts. In fact, the promastigote stage mainly

depends on oxidative phosphorylation for ATP synthesis,66 whereas the amastigote stage has

a higher dependence on glycolysis.41

1.1.5.1.1. Carbon metabolism

Leishmania constitutively expresses most of the main enzymes involved in the carbon

metabolism throughout its life cycle, including those belonging to catabolic pathways for

Page 41: Protein-kinase inhibitors as potential leishmanicidal drugs

26

glucose, amino acids, and fatty acids. It is auxotrophic for purines, vitamins, heme and some

amino acids. Glucose is the preferred carbon source. However, the carbon metabolism is

reprogrammed in the amastigote stage for intracellular growth and survival in the vertebrate

host.65,67,68

The carbon metabolism of the Leishmania promastigote is characterized by a surplus

uptake of glucose, preferentially catabolized via glycolysis and succinate fermentation in the

glycosomes. End-products of glycolysis and succinate fermentation are further catabolized

via the tricarboxylic acid (TCA) in the mitochondrion, where CO2, reducing agents (NADH

and FADH2) and anabolic precursors are produced.67 NADH and FADH2 generated in the

TCA cycle are used to replenish reducing agents at the inner membrane of the

mitochondrion, where they are required for ATP synthesis through oxidative

phosphorylation.66,69 The excess of glucose is managed through an overflow metabolism that

secretes partially oxidized glucose subproducts and avoids overload of the TCA cycle, thus

preventing intracellular stress. Leishmania promastigotes also utilize non-essential amino

acids (aspartate, alanine, and glutamate) besides glucose as a carbon source67,68 (Figure 3,

panel A).

In contrast, the amastigote stage shows a semi-quiescent state with a stringent

metabolism. These metabolic changes are associated with differentiation signals such as

temperature and pH, and not dependent on nutrient limitations in the phagolysosome.68 It is

surmised that the stringent metabolism of the Leishmania amastigote confers resistance

against nutrient depravation and other stresses (oxidative, nitrosative, pH, temperature)

within the parasitophorous vacuole, thus being essential for intracellular survival and

virulence.65 Glucose and amino acid uptake are significantly lower in comparison to the

promastigote stage, but both carbon sources are utilized much more efficiently, showing

scarce overflow metabolism.68 Major energy-consuming processes are specifically repressed

and the flux through the TCA is decreased to avoid overproduction of endogenous ROS due

to electron leaks from the respiratory chain with a high mitochondrial NADH/NAD+ ratio.70

Instead of amino acids, amastigotes resort to fatty acid oxidation as an additional carbon

source besides glucose. Moreover, the metabolic intermediaries of the TCA cycle for

glutamate, glutamine and aspartate biosynthesis are replenished through β-oxidation of fatty

acids65 (Figure 3, panel B).

Page 42: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

27

Figure 3. Central carbon metabolism of promastigotes (A) and amastigotes (B) of L. mexicana.68

Major carbon sources and overflow metabolites are in blue and green boxes respectively. Fluxes through

dotted pathways are down-regulated relative to the opposite stage. Abbreviations: αKG, α-ketoglutarate;

AcCoA, acetyl-CoA; Cit, citrate; Fum, fumarate; FFAA, fatty acid; Glc6P, glucose 6-phosphate; G3P,

glyceraldehyde 3-phosphate; Mal, malate; OAA, oxaloacetate; Ac, acetate; PEP, phosphoenolpyruvate;

PPP, pentose phosphate pathway; Pyr, pyruvate; SCoA, succinyl-CoA; Suc, succinate.

1.1.5.1.2. Oxidative phosphorylation

The mitochondrion of Leishmania promastigotes has a fully functional TCA cycle68

and a regular eukaryotic respiratory chain in their inner mitochondrial membrane, with four

electron transport complexes and a mitochondrial F0/F1 ATP-synthase. Oxidative

phosphorylation takes place due to the coupled activities of these elements: during electron

transport, complexes III and IV generate a transmembrane proton gradient that F0/F1

ATP-synthase (also referred to as complex V) uses for ATP synthesis63 (Figure 4). Although

it is present in the respiratory chain of Leishmania,41 complex I of trypanosomatids lacks

subunits essential for proton pumping, so it is thought that complex I does not play ―or

barely plays― a role in energy generation, and that the main function of this complex is to

maintain a balanced mitochondrial NADH/NAD+ ratio.71 Consequently, the main electron

donor in Leishmania promastigotes is succinate, as substrate for complex II.63 Arguably, a

mitochondrial NADH-dependent fumarate reductase (FRD) present in Leishmania may be

Page 43: Protein-kinase inhibitors as potential leishmanicidal drugs

28

considered an alternative to complex I, as it oxidizes NADH (the natural substrate for

complex I) using fumarate as an electron acceptor to generate succinate, which is further

oxidized by complex II.72

Figure 4. Respiratory chain of Leishmania promastigotes. It consists of complex I (NADH-quinone

oxidoreductase, E.C. 7.1.1.2), complex II (succinate-ubiquinone oxidoreductase, E.C. 1.3.5.1), complex

III (ubiquinol-cytochrome c3 oxidoreductase, E.C. 7.1.1.8) and complex IV (cytochrome c2 oxidase, E.C.

1.9.3.1). Ubiquinone (Coenzyme Q; CoQ) and cytochrome c (CytC) are mobile electron carriers between

the complexes. Complex V is an ATP-synthase that generates mitochondrial ATP using the

transmembrane proton gradient generated by the electron transport chain.63 FRD: fumarate reductase.

Otherwise, as previously mentioned, the amastigote stage of Leishmania adopts a

stringent metabolism that entails reduction of highly energetic cellular processes with

avoidance of oxidative stress.65 Since a surplus of glucose increases mitochondrial NADH

and ROS production, oxidative phosphorylation and hexose uptake are reduced, although

hexose uptake remains essential for survival.73,74 Switching from oxidative phosphorylation

to glycolysis as an energy source is a thriving mechanism shared with other intracellular

pathogens to decrease their susceptibility to exogenous and endogenous RNS/ROS.75

Chakraborty et al.41 observed a downregulation of mitochondrial electron transport in

L. donovani amastigotes with strong reduction of the respiration. Moreover, fumarate served

Page 44: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

29

as an electron sink in the anaerobic metabolism of the parasite, resulting in excretion of the

succinate excess generated. Interestingly, the parasite may couple succinate and proton

pumping as a metabolic energy source for nutrient and ion transport. Plasma membrane

electron transport has been described before in L. donovani promastigotes.76,77

1.1.5.1.3. Other metabolic peculiarities: Folate synthesis, ergosterol and trypanothione

Trypanosomatids also have specific metabolic differences with respect to their hosts.

Folate metabolism: While mammalian cells are auxotrophic for folic acid,

trypanosomatid parasites are capable of de novo folate synthesis. Folates work as

cofactors in the metabolism of amino acids, DNA and RNA synthesis and other essential

metabolic pathways. Interestingly, folate is produced from pteridines, which are

synthesized by insects and mammals. Hence, Leishmania salvages pteridines for folate

biosynthesis from their hosts.78

Ergosterol: Similar to fungi, the major sterol in the plasma membrane of trypanosomatid

protozoa is ergosterol, as opposed to cholesterol in mammals.79

Trypanothione: It is a low molecular mass thiol formed by two molecules of glutathione

conjugated to a single molecule of spermidine. Similar to glutathione in mammalian

cells, trypanothione governs the redox metabolism of Leishmania. It works as an electron

donor for the reduction of peroxides and the formation of deoxyribonucleotides, among

other redox tasks. Therefore, it plays an important role in the antioxidant defences of

intracellular Leishmania against nitrosative and oxidative stresses from the macrophage.

Trypanothione is also involved in the defence against some heavy metals and

xenobiotics, and is essential for the glyoxalase system, which removes toxic and

mutagenic intermediates.80,81

1.1.6. Signal transduction in Leishmania

Signal transduction enables cellular responses to external and internal stimuli,

allowing the cell to adapt to changes and maintain intracellular homeostasis. In parasites

such as Leishmania, the importance of signal transduction is emphasized by the drastic

Page 45: Protein-kinase inhibitors as potential leishmanicidal drugs

30

environmental changes undergone by the parasite throughout its life cycle. Receptors for

these stimuli initiate signal cascades that trigger and regulate expression of certain genes,

activity of specific proteins, and/or changes in cell cycle progression involved in parasite

differentiation, survival, and growth. Members of signalling pathways in eukaryotes, such

as adenylate cyclases, phosphodiesterases, protein kinases and phosphatases among others,

are present in Leishmania.82

Adenylate cyclases (ACs) and phosphodiesterases mediate cAMP-dependent

signalling cascades by regulation of the cAMP intracellular levels through cAMP synthesis

and hydrolysis, respectively. Interestingly, in trypanosomatids it has been suggested that

some ACs play an essential role as signalling receptors due to their extracellular N-terminal

end, which is thought to contain a binding domain for an external ligand. Thus, it is surmised

that ACs in trypanosomatids compensate the lack of G-protein-coupled receptors (GPCRs)

found in other eukaryotes.

Protein kinases (PKs) and protein phosphatases regulate the phosphorylated state of

target proteins or peptides by phosphorylation and dephosphorylation, respectively. Protein

kinases are one of the largest protein families for most eukaryotic organisms, underlying

their relevance in the regulation of cellular growth and survival. Eukaryotic PKs are

characterized by a long-conserved domain (eukaryotic kinase domain) composed of 11

subdomains and a specific three-dimensional structure. They are largely distributed into

three categories according to the amino acid targeted for phosphorylation: Ser/Thr, Tyr, and

dual-specificity kinases. More specifically, eukaryotic PKs are additionally grouped by the

Manning classification into nine groups (Table 2) following phylogenetic criteria (Figure

5).83 Alternatively, protein phosphorylating enzymes devoid of the eukaryotic kinase domain

are labelled as atypical PKs and PK-like.84

Page 46: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

31

Table 2. Eukaryotic protein kinases groups according to Manning.83

Eukaryotic protein

kinase groups Name description

AGC Protein kinase A, G and C families

CAMK Ca2+/Calmodulin-dependent kinase group

CMGC CDK, MAPK, GSK, and CLK families

STE Homologs of yeast sterile 7, 11, and 20

CK1 Casein Kinase 1 group

TK Tyrosine kinase group

TKL Tyrosine kinase-like group

RGC Receptor guanylate cyclase group

“Other” Kinase families that do not fit elsewhere

1.1.6.1. The kinome of Leishmania

Comparative studies of the kinomes of L. major, Trypanosoma brucei, T. cruzi,

L. infantum and L. braziliensis identified 196, 176, 190, 224 and 221 putative protein kinases

respectively, and observed some remarkable peculiarities of the trypanosomatid

kinomes:85,86

The AGC and the CAMK groups are relatively poorly represented (13.8-15.6% of their

kinome), whereas the CMGC and the STE groups are fairly well represented (more than

40% of their PK repertoire), as compared to these paired PK groups in the human kinome

(26.4% and 20.8% respectively).83

Trypanosomatidae parasites are apparently devoid of protein kinases G or

cGMP-dependent (PKGs).82

Although trypanosomatids do not have any tyrosine kinase or tyrosine kinase-like

groups, it is well known that phosphorylation on tyrosine occurs. Parsons et al.85

Page 47: Protein-kinase inhibitors as potential leishmanicidal drugs

32

proposed that atypical TKs and dual-specificity kinases encoded in the trypanosomatid

genomes carry out this task.

Figure 5. Phylogenetic classification and distribution of human protein kinases.87

Page 48: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

33

In addition, Borba et al.86 identified 157 Leishmania-specific protein kinases after

comparing the kinomes of L. infantum, L. braziliensis and humans. These

Leishmania-specific PKs were further categorized in functional groups; almost one third

(31%) of these enzymes were assigned to signal transduction, while the remainder were

distributed into cell growth and death processes (36%), and a variety of other functions

(metabolism, environmental adaptation, cell motility, morphogenesis…). Moreover,

phosphoproteomic studies in Leishmania endorsed the regulation of PKs as essential for

promastigote-to-amastigote differentiation, since phosphorylation of most proteins in the

parasite occurs in a stage-specific manner (promastigote or amastigote) rather than

constitutively (in both stages).88,89

Therefore, protein kinases are appealing targets for Leishmania chemotherapy due to

their variety of essential roles in the parasite. In fact, some of them have already been

validated as pharmacological targets for the development of new leishmanicidal drugs

(Table 3), although the roles of most PKs identified in Leishmania are still unclear.

Page 49: Protein-kinase inhibitors as potential leishmanicidal drugs

3

4 34 Table 3. Eukaryotic protein kinases (ePKs) described in the kinomes of Leishmania spp.85,86,90

ePK

groups

PK families in

Leishmania Function

Validated target

in Leishmania

AGC

Protein kinase A or

cAMP-dependent (PKA)

Phosphorylation of downstream protein targets in response to an increase in cAMP

intracellular levels.91 In Leishmania spp., PKA is almost exclusively active in the metacyclic

promastigote, hence it is likely relevant for the primary invasion of the macrophage.82

No

Protein kinase C (PKC) PKC-like activity in Leishmania has been reported,92 although its encoding gene has not been

identified in any of the trypanosomatid genomes analysed so far. No

Protein kinase B (PKB)

or Akt

An Akt-like kinase is upregulated in Leishmania under stress conditions, such as nutrient

deficiency and heat shock.93,94 No

Nuclear DBF2-related

kinases (NDR)

Kinome analyses support the existence of NDR kinases in Leishmania, but there is no further

knowledge about them. Two NDR kinases in T. brucei, PK50 and PK53, are essential for a

successful cytokinesis of its bloodstream stage.95

No

Phosphoinositide-

dependent protein kinase

1 (PDK1)

Involved in the regulation of relevant processes of metabolism, growth, proliferation and

survival, driven by lipid signalling in higher eukaryotic organisms.96,97 There is no further

information on specific PDK1 kinases of Leishmania.

No

Ribosomal S6 kinase

(RSK)

Downstream in the mTOR pathway in mammals. S6K1 is responsible for the phosphorylation

of the ribosomal protein S6, with increase of protein synthesis and proliferation ensuing.98

There are no reports regarding the functionality of any RSK of Leishmania.

No

Page 50: Protein-kinase inhibitors as potential leishmanicidal drugs

35

Intro

ductio

n

3

5

Table 3 (continued). Eukaryotic protein kinases (ePKs) described in the kinomes of Leishmania spp.85,86,90

ePK

groups

PK families in

Leishmania Function

Validated target

in Leishmania

CAMK

CAMK1

Member of the calmodulin-dependent protein kinase cascade. The gene of CAMK1 has been

identified in the Leishmania kinome without further notes on its functionality. CAMKs in

other eukaryotic organisms are involved in gene transcription, translation, ion channel

regulation, and cell death processes.99

No

CAMK-like kinase

(CAMKL)

Genes for CAMK-like kinases have been described in Leishmania but are yet to be studied.

A CAMK-like protein of the Apicomplexa Toxoplasma gondii is surmised to be a virulence

factor through the phosphorylation of host proteins.100

No

Ca2+-dependent protein

kinase (CDPK)

CDPKs of Leishmania have not been studied. Their role in other Protozoa, such as the

apicomplexan Plasmodium falciparum, is the regulation of cytosolic Ca2+.101 No

CMGC

Cyclin-dependent kinase

(CDK) or cdc2-related

kinase (CRK)

Crucial for the control of the cell cycle of the parasite. CRK1 and CRK3 are essential for G1/S

transition. CRK3 is also key for G1 and G2/M phase progression together with mitotic cyclins

CYC2 and CYC6, respectively.102,103 CRK12 is essential for growth and survival although its

precise role is unclear.104

CRK1, CRK386,105

CRK12104

Mitogen-activated

protein kinase (MAPK)

MAPKs of Leishmania spp. are involved in flagellar growth regulation (LMAPK3,

LmxMAPK9, LmxMAPK13, LmxMAPK14), stage-differentiation (LmxMAPK1,

LmxMAPK2, LMAPK3, LmxMAPK4), parasite growth (LmjMAPK7), morphology

(LmxMAPK6), and intracellular survival (LmxMAPK1, LmxMAPK2).106,107 LdMAPK1 also

regulates the activity of heat shock proteins.108

MAPK3107

MAPK4109

Page 51: Protein-kinase inhibitors as potential leishmanicidal drugs

36 Table 3 (continued). Eukaryotic protein kinases (ePKs) described in the kinomes of Leishmania spp.85,86,90

ePK

groups

PK families in

Leishmania Function

Validated target in

Leishmania

CMGC

Glycogen synthase

kinase-3 (GSK-3)

Mammalian GSK-3 is a multi-task signalling protein involved in energy metabolism,

proliferation, differentiation, immunity, and cell death and survival.110

Leishmania has short and long GSK-3 isoforms, and the former is crucial for parasite cell

growth.111

LdGSK-3111

Cell division control

(CDC)-like kinase

(CLK),

SR-rich protein kinase

(SRPK),

Dual-specificity tyrosine-

regulated kinase (DYRK)

In higher eukaryotes, CLKs and SRPKs are involved in RNA processing and splicing by

regulating the cellular distribution and activity of SR proteins,112 whereas DYRK1 is known

to pre-phosphorylate substrates of GSK-3.113

Additionally, as dual-specific protein kinases, CLKs and DYRKs are involved in tyrosine

phosphorylation in trypanosomatids.86

No

Casein Kinase 2 (CK2)

CK2 in Leishmania is secreted, but also distributed within the cytoplasm and across the

external surface of the plasma membrane of the parasite. It is involved in promastigote

proliferation and infectivity.114

No

ROS-cross-hybridizing

kinase (RCK)

RCKs have been described as MAPK-related in other protozoa and higher organisms. They

are involved in cilia length and intraflagellar transport of protein complexes along the

axoneme.115-117 Thus, it may be inferred that RCKs in Leishmania are likely involved in

processes of the flagellum, like other MAPKs.

No

Page 52: Protein-kinase inhibitors as potential leishmanicidal drugs

Intro

ductio

n

37

Table 3 (continued). Eukaryotic protein kinases (ePKs) described in the kinomes of Leishmania spp.85,86,90

ePK

groups

PK families in

Leishmania Function

Validated target

in Leishmania

CK1

CK1

CK1 kinases are involved in the regulation of the cell cycle, cellular transport, receptor

signalling, transcription, DNA repair and apoptosis.

In Leishmania, CK1 isoform 2 (CK1.2) is secreted and phosphorylates host substrates,

promoting infection of the host cell.118,119

CK1.2118

Tau tubulin kinase

(TTBK)

TTBKs of Leishmania have not been studied. In other organisms TTBKs phosphorylate tau

microtubule-associated protein and tubulin.120 No

STE

STE7/MAPKK

MAPKKs are dual-specific86 protein kinases homologs of yeast sterile 7 (STE7). They

phosphorylate MAPKs; as such, they participate in the regulation of parasite differentiation

and the maintenance of the flagellum length,85 as well as tyrosine phosphorylation.86

No

STE11/MAPKKK

The MAPKKKs, homologs of yeast sterile 11 (STE11), are abundant in trypanosomatids.

They activate MAPKKs by phosphorylation upstream the MAPK signalling pathways.85,121 In

T. brucei, STE11 was associated with cell growth and stage differentiation.122

No

STE20/MAPKKKK

Homologs of the yeast sterile 20 (STE20) are rare in trypanosomatids.85 They can function as

scaffolds for pathway assembly and MAPKKK self-activation; otherwise they can directly

phosphorylate MAPKKKs.121 Although Leishmania STE20 has not been studied, STE20 of

T. brucei is essential for NDR kinase activation.95

No

Page 53: Protein-kinase inhibitors as potential leishmanicidal drugs

3

8

38 Table 3 (continued). Eukaryotic protein kinases (ePKs) described in the kinomes of Leishmania spp.85,86,90

ePK

groups

PK families in

Leishmania Function

Validated target

in Leishmania

“Other”

Aurora kinase (AIRK) Aurora kinases in Leishmania are surmised to regulate chromosome segregation and

cytokinesis, as in other eukaryotic cells.123

Aurora kinase 1

(AIRK)123

Polo-like kinase (PLK)

PLK genes were found in Leishmania, but their functionality has not been described yet.

In T. brucei, PLKs are essential during mitosis, playing a role in kDNA segregation, basal

body duplication and cytokinesis.124

No

Never in mitosis gene A

(NimA)-related kinase

(NEK)

NEKs are present in Leishmania genomes but have not been studied in detail.

RDK2 is a NEK essential for parasite proliferation in T. brucei,122 accordingly to the role

played by NEKs in cell cycle progression and cytoskeletal functions.85,123

No

Wee1-like protein kinase

(WEE1)

No specific studies on Leishmania WEE1 have been conducted. Nevertheless, WEE1 kinase

is known to participate in tyrosine phosphorylation86 and it is involved in cell division,85 as it

has been described for T. brucei.125

No

Page 54: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

39

1.1.7. State of the art of the chemotherapy for leishmaniasis

As previously stated, the only current treatment available for leishmaniasis is

chemotherapy.126,127 However, the existing drug repertoire is scarce, with downsides such as

toxicity, high cost, parenteral administration and emergence of resistance.128 Two

approaches are being used to fix this scenario. In a long-term strategy, the pipeline for

leishmanicidal drugs is fed with new compounds to broaden chemotherapeutic options.127 In

a short-term range, combination therapy of current antileishmanial drugs successfully

reduces dosing, treatment duration, toxicity and drug resistance.25

Current drugs for leishmaniasis chemotherapy approved by the FDA129 encompasses

pentavalent antimonials, amphotericin B, miltefosine and paromomycin as first-line options,

and pentamidine as a second-line alternative (Figure 6, Table 4).

Figure 6. Chemical structures of the current drugs approved for the treatment of human

leishmaniasis.

Page 55: Protein-kinase inhibitors as potential leishmanicidal drugs

40 Table 4. Main current drugs available for chemotherapeutic treatment of human leishmaniasis. (i.v.: intravenously, i.m.: intramuscularly)

Leishmanicidal drugs Adm. route Mechanism of action Side effects and downsides Resistance References

PENTAVALENT

ANTIMONIALS

Meglumine antimoniate

(MA)

Sodium stibogluconate (SSG)

i.v.

or

i.m.

Interferes with the redox

metabolism by targeting the

trypanothione system of the

Leishmania amastigote.

Severe side effects such as

hepatic and renal toxicity and

cardiac abnormalities. Requires

extended hospitalization with

higher costs.

Emerging resistance

especially in the

Indian subcontinent.

128,130

AMPHOTERICIN B

Deoxycholate

(AmBd)

i.m

Causes plasma membrane

permeabilization by selective

binding to ergosterol of the

parasite membrane.

Highly toxic, frequently causes

electrolyte abnormalities, ne-

phrotoxicity and anaemia.

There are sparse

cases of clinical

AmB resistance.

127,128 AMPHOTERICIN B

liposomal formulation

(LAmB)

(Ambisome)

Significantly lower toxicity than

AmBd, but its implementation is

restricted due to its higher cost

and cold-chain transport require-

ment.

PAROMOMYCIN

i.m. for VL,

topical for

CL

Selectively interferes in

Leishmania ribosomal activity,

inhibiting protein synthesis.

Affects mitochondrial membrane

potential causing inhibition of

respiration.

Injection-site pain, hepato- and

ototoxicity, and renal dys-

function.

Laboratory-induced

resistance has been

described.

127,128,131

Page 56: Protein-kinase inhibitors as potential leishmanicidal drugs

Intro

ductio

n

41

Table 4 (continued). Main current drugs available for chemotherapeutic treatment of leishmaniasis. (i.v.: intravenously, i.m.: intramuscularly)

Leishmanicidal drug Adm. route Mechanism of action Side effects and downsides Resistance References

MILTEFOSINE Oral

Modifies cell membrane com-

position by interfering with

phosphatidylcholine biosynthesis.

Significantly reduces mitochon-

drial membrane potential and

cytochrome c oxidase activity.

It may promote apoptosis through

Akt inhibition.

Immunomodulatory activity in

macrophages.

Teratogenic. Causes gastro-

intestinal discomfort.

Broad range of sensitivity among

Leishmania species, with low

efficacy in HIV co-infection

cases.

Its long half-life favours drug

resistance.

Easily induced in

vitro.

Growing reports of

resistant parasites in

clinical settings.

127,132,133

PENTAMIDINE

i.v.

or

i.m.

The mitochondrion is likely an

important target, but more targets

are suspected.

Severe adverse effects, from

cardiotoxicity to several

metabolic and electrolyte

disturbances (hypo- or

hyperglycaemia, type 1 diabetes

mellitus, hyperkalaemia,

hypocalcaemia…)

Unresponsiveness in

Indian patients.

49,127,131

Page 57: Protein-kinase inhibitors as potential leishmanicidal drugs

42

Additional approaches in the search for new leishmanicidal treatments involve drug

repositioning as well as target-based and phenotypic screenings.134

Target-based methods aim to find new inhibitors against a well-defined target, e.g. a

specific enzyme. Identified hits are compounds with a known mechanism of action that

need further validation in a cellular model.

Phenotypic approaches rely on the screening of compounds directly against the parasite.

Hit compounds inhibit cellular growth, but further target deconvolution is desirable in

order to unveil their mechanism of action.

In drug repositioning, compounds initially developed against a specific pathology are

assayed for leishmanicidal activity. An important asset of this method is that most of the

toxicological and preclinical data of the compounds are already available.135 Some of the

current drugs used in the treatment of leishmaniasis are, in fact, repositioned drugs:

miltefosine was originally developed as an anticancer drug, paromomycin as a

bactericidal antibiotic and amphotericin B is an antifungal.127

Some years ago, the Drugs for Neglected Diseases initiative (DNDi) initiated a search

for novel leishmanicidal agents and established target product profiles (TPPs) in close

association to academic groups. TPPs define the parameters and criteria of each stage of

drug development: administration route, dosing regimen, safety and tolerability levels, cost

and shelf life.136 Nowadays, different institutions and big pharma companies collaborate with

DNDi in phenotypic and target-based screenings of their chemical libraries.28 In recent years,

five new lead series have reached preclinical stages of development such as the

aminopyridazole DNDI-5561, and even clinical phase I like the oxaborole DNDI-6148, the

nitroimidazole DNDI-0690 or the pyrazolopyrimidine GSK3186899/DDD853651

(Figure 7). Although their mechanism of action is not fully understood, some clues about the

biochemical targets in Leishmania have been envisaged.137

Page 58: Protein-kinase inhibitors as potential leishmanicidal drugs

Introduction

43

Figure 7. Leishmanicidal compounds of the DNDi pipeline that have reached Phase I.

Page 59: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 60: Protein-kinase inhibitors as potential leishmanicidal drugs

OBJECTIVES

Page 61: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 62: Protein-kinase inhibitors as potential leishmanicidal drugs

Objectives

47

2. Objectives

Due to the urgent need for new effective treatments for leishmaniasis, the main

objective of this PhD thesis is the discovery of new leishmanicidal agents with a known

mechanism of action. For this purpose, we have evaluated different sets of compounds under

a target-based approach focused on the inhibition of the short isoform of GSK-3 of

Leishmania, a validated drug target against the parasite, in combination with a phenotypic

screening on axenic promastigotes and amastigotes. Moreover, as energy metabolism is a

reliable parameter of cell viability, bioenergetic parameters were studied as a beacon to

assess the physiological homeostasis of the parasite and explore off-target effects.

The specific objectives set for this purpose were as follows:

1.- To select human protein kinase inhibitors (hPKIs) with representative scaffolds

from our in-house chemical library as potential LdGSK-3s inhibitors.

2.- To select virtual screening hits that putatively bind to pharmacologically validated

kinases in Leishmania.

3.- To test the selected compounds against LdGSK-3s in a target-based approach.

4.- To assess the leishmanicidal activity and cytotoxicity on murine macrophages of

the selected compounds under a phenotypic screening strategy.

5.- To establish structure-activity correlations for those PKIs with leishmanicidal

activity and LdGSK-3s inhibition, identified as hits.

6.- To evaluate the capability of our hits to decrease the parasite load in peritoneal

murine macrophage infections.

7.- To test inhibition of the energy metabolism of Leishmania as an additional

off-target mechanism of our hits.

Page 63: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 64: Protein-kinase inhibitors as potential leishmanicidal drugs

MATERIALS AND METHODS

Page 65: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 66: Protein-kinase inhibitors as potential leishmanicidal drugs

Materials and Methods

51

3. Materials and Methods

3.1. Chemical compounds, media, reagents and laboratory equipment

This work reports the evaluation of 136 heterocyclic compounds with low molecular

weight (MW <500) and purity 95 % by HPLC from our in-house chemical library, named

MBC chemical library.138 The tested compounds differ in their chemical scaffolds and

substitution patterns. In this work, each compound was labelled with a number from 1 to 136

(Tables 7, 8, 10, 11, 12 and 13).

Reagents, materials, media and equipment used in cell cultures, buffers and assays

were purchased from Sigma-Aldrich unless stated otherwise.

3.2. Cell cultures and cell harvesting

The Leishmania cell lines used in this project were Leishmania infantum JPC

(MCAN/ES/98/LLM-722) and 3-LUC L. donovani (MHOM/SD/00/1S-2D) promastigotes,

as well as L. pifanoi (MHOM/VE/60/Ltrod) and mCherry-L. pifanoi axenic amastigotes.

Elicited murine peritoneal macrophages (MPM) from Balb/C mice (9 to 14-week-old males)

were also used.

L. infantum promastigotes were grown at 26 ºC in RPMI 1640 medium supplemented

with 10% heat inactivated fetal calf serum (RPMI-HIFCS; Table 5). L. donovani 3-LUC

promastigotes, a modified strain of L. donovani promastigotes that expresses a mutated

Phothinus pyralis luciferase gene, were grown in the same conditions with addition of

30 µg/mL geneticin in the medium.139

L. pifanoi axenic amastigotes were grown at 32 ºC in M199 supplemented with

20% HIFCS (M199-HIFCS; Table 5). mCherry-L. pifanoi axenic amastigotes, a strain

Page 67: Protein-kinase inhibitors as potential leishmanicidal drugs

52

transfected with an expression vector encoding the gene for the red fluorescent protein

mCherry, were cultured in the same conditions plus 50 µg/mL blasticidin hydrochloride in

the medium.

Table 5. Media and buffer composition.

Medium/Buffer Composition

RPMI-HIFCS

RPMI 1640 medium (Gibco) supplemented with 10% Heat Inactivated Fetal

Calf Serum (HIFCS), 5 mM HEPES (Biowest), 1.7 mM NaHCO3, 2 mM

L-glutamine, 20 U/mL unicilin (ERN, S.A.) and 24 µg/mL gentamicin

(NORMON, S.A.), pH 6.8-6.9.

M199-HIFCS

M199 medium (Gibco) supplemented with 20% HIFCS, 0.5% trypticase

peptone (BD Biosciences), 13.9 mM D-glucose, 76.7 µM hemin, 5.1 mM

glutamine and 40 µg/mL gentamicin; pH 7-7.2.

DMEM-HIFCS DMEM medium (Gibco) supplemented with 10% HIFCS, 2 mM

L-glutamine and 100 U/mL penicillin-streptomycin.

HBSS-Glc

Hank’s Balanced Salt Solution (137 mM NaCl, 5.4 mM KCl, 0.4 mM

KH2PO4, 4.2 mM NaHCO3, 0.4 mM Na2HPO4; pH 7.2-7.3) supplemented

with 1% D-glucose.

Leishmania parasites were harvested at late exponential growth phase (~107 cells/mL)

by 10 min centrifugation (10 min, 1,600 ×g, 4 ºC).

MPM harvesting was carried out in mice by elicitation with thioglycollate

(intraperitoneal injection, 1 mL 10% thioglycollate, 72 h). Elicited macrophages were

extracted by peritoneal wash (×2) with 10 mL HBSS-Glc (Table 5) using 25G and 21G

needles (BD MicrolanceTM), followed by centrifugation (10 min, 400 ×g, 4 ºC).

3.3. Leishmanicidal activity of the compounds on axenic parasites

The 136 compounds from Sets 1 to 4 and Sets 1.1 and 2.1 were tested for

leishmanicidal activity by phenotypic screening on axenic promastigotes and amastigotes

Page 68: Protein-kinase inhibitors as potential leishmanicidal drugs

Materials and Methods

53

through a colorimetric assay based on MTT reduction to formazan. Under these conditions,

inhibition of MTT reduction accounts for the inhibition of parasite proliferation.

Parasites resuspended in their respective growth medium were aliquoted into

96-microwell plates (2×106 parasites/mL, 200 µL/well) and incubated with the

corresponding compound (72 h at 24 ºC for L. infantum promastigotes; 96 h at 32 ºC for

L. pifanoi axenic amastigotes).

In order to avoid interference in the readout of formazan, L. infantum promastigotes

were prepared in RPMI-HIFCS using RPMI-1640 without phenol red (Gibco)

(RPMIØRED-HIFCS). L. pifanoi axenic amastigotes were prepared in M199-HIFCS, and

removal of the medium was required due to colorimetric interference of hemin. For this, the

content of each well plus an additional wash with 200 µL HBSS-Glc were collected in

1.5 mL Eppendorf tubes. Amastigotes were further diluted with additional 800 µL

HBSS-Glc. After centrifugation (4 min, 13,000 ×g, 4 ºC), pellets were resuspended in

200 µL HBSS-Glc and transferred back into a 96-microwell plate.

For all axenic parasites, MTT was added to each well (0.5 mg/mL, final concentration).

MTT reduction by viable parasites was stopped within 2 h of incubation with 10% SDS

(50 µL/well). Afterwards, the solubilized formazan was read at 595 nm in a 680 BioRad

Microplate-reader.

Each sample was made by triplicate and assays were repeated at least twice. The IC50s

(concentration required for 50% inhibition in vitro) of leishmanicidal compounds were

calculated with SigmaPlot v11.0.

3.4. Macrophage cytotoxicity of leishmanicidal compounds

MPMs in DMEM-HIFCS (Table 5) were seeded in 96-well plates (1×105 cells/well).

Cells were allowed to adhere to the bottom of the wells (24 h at 37 ºC, 5% CO2), followed

by removal of the medium and addition of compounds in RPMIØRED-HIFCS. After 48 h

of incubation, MTT reduction was carried out as stated in Section 3.3.

Page 69: Protein-kinase inhibitors as potential leishmanicidal drugs

54

Experiment design and IC50 calculations were conducted as previously described (see

Section 3.3). The Selectivity Index (SI) of each compound was defined as the ratio IC50 axenic

amastigote/IC50 macrophage.

3.5. Leishmanicidal activity against intracellular amastigotes

MPMs resuspended in DMEM-HIFCS (Table 5) were transferred into a 24-microwell

plate (2×105 cells/well) with a sterile round 22 mm diameter coverslide at the bottom of each

well. Cells were allowed to adhere to the glass surface of the coverslides for 24 h (37 ºC,

5% CO2).

Afterwards, the medium was replaced with a 0.5 mL suspension of mCherry-L. pifanoi

axenic amastigotes (6×105 cells/well; 3:1 ratio parasite:macrophage) in RPMI-HIFCS.

Phagocytosis was allowed to proceed for 4 h at 32 ºC. Non-phagocytosed amastigotes were

removed by gentle washing of the wells (×3) with 1 mL of HBSS at 32 ºC. Next, 1 mL/well

RMPI-HIFCS was added and the infection was left to proceed for 24 h at 32 ºC. Finally, the

infected macrophages were incubated with 1 mL of compound at its corresponding

concentration in new RPMI-HIFCS for 16-20 h, at 32 ºC. The parasite index (number of

amastigotes per macrophage) was evaluated with optical and fluorescence microscopy

(Leica DIALUX 20). At least six different fields gathering up to 100 macrophages were

counted in each preparation.

Each sample was made by duplicate and experiments were repeated at least twice.

3.6. Extraction and purification of LdGSK-3s

E. coli BL21(D3) was transfected by electroporation with the expression vector

pET28a(+) containing the gene of the short isoform of L. donovani GSK-3 (LdGSK-3s),

kindly donated by Dr. Smirlis (Hellenic Pasteur Institute).111 The LdGSK-3s gene was

C-terminally fused to a poly-His tag and placed at the restriction site of the polylinker Nco I

and Xho I under the T7lac promoter. The process of expression, extraction and purification

Page 70: Protein-kinase inhibitors as potential leishmanicidal drugs

Materials and Methods

55

of LdGSK-3s is compiled in Figure 8, whereas the buffers used in each stage of the process

are described in Table 6. Fractions and samples of each purification step were analysed for

their protein pattern by SDS-PAGE electrophoresis on a 10% polyacrylamide gel. Fractions

with high purity and protein content of LdGSK-3s were pooled and dialysed with an Amicon

cell filtration (30 kDa cut-off membrane) using Kinase PBS as washing buffer. Lastly,

LdGSK-3s was quantified after ultracentrifugation (147,000 ×g, 1 h, 4 ºC) and filtration

through a 0.22 μm nitrocellulose filter, and preserved at -80 ºC.

Figure 8. Flowchart of the expression, extraction and purification of LdGSK-3s. French Press:

Thermo Scientfic; Ni2+ column: HisTrapTM FF crude 5 mL (Merck, Ref. 17-5286-01).

Page 71: Protein-kinase inhibitors as potential leishmanicidal drugs

56

Table 6. Buffers used for LdGSK-3s purification, dialysis, and activity evaluation.

Buffer Composition

Kinase PBS

150 mM NaCl, 1.5 mM H2KPO4, 2.7 mM KCl, 8.3 mM HNa2PO4, 60 mM

β-glycerophosphate disodium, 1 mM Na3VO3, 1 mM NaF, 1 mM disodium

phenyl phosphate; pH 7.5

Lysis/Binding

Buffer

Kinase PBS with 10mM imidazole, plus protease inhibitors cocktail (Roche

Ref. 1697498); pH 7.5

Washing Buffer

Kinase PBS plus additional 150 mM NaCl (300 mM NaCl final concentration),

30 mM imidazole, 1% Triton X-100 (TX-100), and protease inhibitors cocktail;

pH 7.5

Elution Buffer Kinase PBS with 300 mM imidazole plus protease inhibitors cocktail; pH 7.5

Kinase Assay Buffer 50 mM HEPES, 1 mM EGTA, 1 mM EDTA, 15 mM magnesium acetate, 0.1

mg/mL BSA; pH 7.5

3.7. Measurement of LdGSK-3s activity

Enzymatic activity of selected compounds was tested using 20 ng of purified

LdGSK-3s in the presence of 25 μM phosphoglycogen synthase peptide-2 (GS-2; Millipore,

Ref. 12-241) and 1 μM ATP (Sigma-Aldrich, Ref. A7699) in 40 μL Kinase Assay Buffer in

black 96-microwell plates. The reaction proceeded for 30 min at 30 ºC, followed by addition

of 40 μL Kinase Glo reagents (Promega, Ref. V6712). After 10 min of incubation at room

temperature, the remaining ATP from the kinase reaction was measured by an end-point

luminescence readout obtained in a Varioskan Flash microplate reader.

Each sample was made by duplicate and assays were repeated at least twice. The

following controls were included: i) samples with GS-2 and ATP (without LdGSK-3s) to

obtain the maximum luminescent signal, ii) samples with GS-2, ATP and LdGSK-3s to

obtain the minimum luminescent signal due to maximal LdGSK-3s activity, iii) samples with

GS-2, ATP, LdGSK-3s and a commercial inhibitor (10 μM indirubin-3’-monoxime-5-

sulphonic acid)140 to demonstrate the specific activity of LdGSK-3s. The IC50s against

Page 72: Protein-kinase inhibitors as potential leishmanicidal drugs

Materials and Methods

57

LdGSK-3s activity were calculated with the statistical programs included in SigmaPlot

v11.0.

3.8. Bioenergetic assays

3.8.1. In vivo monitoring of intracellular ATP levels

Real-time variation of Leishmania intracellular ATP levels was studied using 3-LUC

L. donovani promastigotes.139 This strain was obtained by electroporation of the parental

strain with a pLEXSY expression vector encoding a cytoplasmic luciferase from Phothinus

pyralis, mutated at the import signal for glycosomes located at the C-terminal tripeptide,

preventing its transport into this organelle.

Promastigotes (22×106 parasites/mL) in HBSS-Glc plus 50 μM DMNPE D-luciferin

(GOLDBIO, USA) were immediately aliquoted into a black 96-microwell plate

(90 μL/well). Luminescence was monitored in a POLARstar Galaxy microplate reader

(BMG LABTECH). Once a steady readout of luminescence was obtained, 10 µL of the

corresponding compounds in HBSS-Glc at 10-fold final concentration were added, and

changes in luminescence were followed up for at least 30 min.

Samples were made by duplicate and assays were repeated at least twice. Positive

controls for plasma membrane permeabilization (0.1% TX-100) and inhibition of oxidative

phosphorylation (1.5 µM 1,4-naphthoquinone) were also included.

3.8.2. Assessment of plasma membrane permeabilization

Leishmania parasites (22×106 parasites/mL) in HBSS-Glc plus 1µM SYTOXTM Green

(Invitrogen) were dispensed into a 96-microwell black plate (90 μL/well). Increase in the

fluorescence of SYTOXTM Green was monitored in a POLARstar Galaxy microplate reader

(λEX = 485 nm, λEM = 520 nm). After an initial reading of the basal fluorescence, 10 µL of

the corresponding compounds at 10-fold final concentration in HBSS-Glc were added.

Page 73: Protein-kinase inhibitors as potential leishmanicidal drugs

58

Variation of the fluorescence was followed up for ~30 min. Afterwards, maximal

permeabilization value for each well (100%) was established by the addition of 10 µL

1% TX-100.141

Samples were prepared by duplicate and assays were repeated at least twice. TX-100

(0.1%, final concentration) was used as a positive control of plasma membrane

permeabilization.

3.8.3. Mitochondrial membrane potential (ΔΨm) of Leishmania

Variation of mitochondrial membrane polarization was evaluated through the

intracellular accumulation of the fluorescent dye rhodamine 123, which is driven by the

electrochemical potential of the mitochondrion. Leishmania parasites in HBSS-Glc

(20×106 parasites/mL, 100 µL final volume) were incubated with the compounds in 1.5 mL

eppendorf tubes for 4 h at the optimal temperature for each parasite (26 ºC for promastigotes,

32 ºC for amastigotes). The medium was removed after centrifugation (4 min, 13,000 ×g,

4 ºC) and the pellet resuspended in 100 µL 0.3 µg/mL rhodamine 123 and incubated for

10 min at room temperature in darkness. After washing the samples with HBSS-Glc to

remove extracellular dye, the pellet was resuspended in 900 µL HBSS-Glc and transferred

to cytometry tubes. Rhodamine 123 fluorescence was measured in a Coulter XL EPICS flow

cytometer (λEX = 488 nm, λEM = 520 nm).141

Samples were prepared by duplicate and assays were repeated at least twice. A

40 min-treatment with 20 mM KCN was used as a depolarization control.

3.8.4. Evaluation of the O2 consumption rate of the parasite

Variation in the O2 consumption rate of L. infantum promastigotes was measured in a

Clark electrode (Hansatech Instruments) at 1×108 cells/mL in 600 μL of Respiration Buffer

(10 mM Tris-HCl, 125 mM sucrose, 65 mM KCl, 1 mM MgCl2, 2.5 mM NaH2PO4, 0.3 mM

EGTA, 5 mM succinic acid; pH 7.2). Compounds were added with a Hamilton pipette at

Page 74: Protein-kinase inhibitors as potential leishmanicidal drugs

Materials and Methods

59

100-fold final concentration to whole parasites, and oxygen consumption was monitored for

8-10 min.

In order to identify the targets of the selected compounds, they, as well as substrates

and inhibitors of the respiratory chain, were also tested on the mitochondrion of Leishmania

promastigotes permeabilized with 60 µM digitonin. Before the addition of the pertinent

treatment, Leishmania mitochondria were supplemented with 100 μM ADP to ensure O2 was

the limiting substrate for respiration.

Each assay was repeated at least twice. Untreated parasites and mitochondrion were

used as a control for basal respiration. Malonate (10 mM) and oligomycin (12.6 μM) were

used as controls for the inhibition of Complexes II and V respectively. FCCP (10 μM) was

used as an uncoupler control. Substrates used included 6.7 mM α-glycerophosphate for

Complex I, and 0.1 mM TMPD-1.7 mM ascorbate for Complex IV.

3.8.5. Assessment of programed cell death induced by the compounds

L. infantum promastigotes (2×106 cells/mL) in RPMIØRED-HIFCS were incubated

with the compounds in 24-well plates (1 mL/well) for 72 h at 26 ºC. Next, samples were

transferred to 1.5 mL Eppendorf tubes and centrifuged (15,700 ×g, 4 min, 4 ºC). Each pellet

was resuspended in 1 mL HBSS-Glc and centrifuged again in order to further wash away

the compounds. Next, the pellets were resuspended in 15 μL of the supernatant. Fixation and

permeabilization of the parasites were carried out by careful addition of 200 μL 70% cold

ethanol to each resuspended pellet and overnight incubation at 4 ºC. Ethanol was removed

by centrifugation and washing of the samples (×2) with 1 mL HBSS-Glc. Finally, each pellet

was resuspended in 500 μL HBSS-Glc with 20 μg/mL propidium iodide (PI) and 3 mg/mL

Ribonuclease A, and incubated in darkness at room temperature in cytometry tubes for

30 min. Fluorescence of PI was measured in a Coulter XL EPICS flow cytometer

(λEX = 488 nm, λEM = 620 nm).

Page 75: Protein-kinase inhibitors as potential leishmanicidal drugs

60

Each assay was repeated at least twice. Untreated cells were used as control for a

standard cell cycle histogram, and miltefosine (15 μM) was used as control for an

apoptosis-inducing respiratory inhibitor.

Page 76: Protein-kinase inhibitors as potential leishmanicidal drugs

RESULTS AND DISCUSSION

Page 77: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 78: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

63

4. Results and discussion

Since no vaccines against human leishmaniasis are commercially available yet, the

treatment against Leishmania exclusively relies on chemotherapy despite its loss of

effectiveness over the last decades due to side effects, high costs, and emergence of

resistance.23-25 Few leads have been recruited in the pipeline for drug discovery against

leishmaniasis,142 but none of them have reached clinical phase II yet, and the mechanism of

action of many remains unveiled. The lack of preclinical candidates renders the pipeline

extremely fragile to cope with the inherent attrition of any drug discovery and development

program. Thus, it is imperative to find new hits.

One of the best approaches to feed the pipeline is the screening of chemical collections

with potential leishmanicidal activity. These screenings can be phenotypic or target-based.

Despite the importance of target identification for later stages in drug development, the

DNDi guidelines do not prioritize target-based screenings over phenotypic ones, prompted

by the urgent need for new hits and the scarce number of validated targets for kinetoplastid

parasites. Thus, the current pipeline for novel antileishmanial drugs is mostly sustained by

hits selected from phenotypic screenings.28 In this work, we focused our interest in new

leishmanicidal hits with a presumed mechanism of action without dismissing the feasibility

of off-side targets. To this end, phenotypic and target-based screenings were concurrently

carried out, with selected compounds from our in-house chemical library assayed for

inhibition of both Leishmania proliferation and Leishmania GSK-3s (LGSK-3s) activity.

GSK-3 is a protein kinase highly conserved throughout eukaryotic organisms, from

yeast and protozoa to plants and mammals. It is a ubiquitous multi-task serine-threonine PK

of the CMGC family. More than 100 substrates for this enzyme have been reported,143 with

over 40 identified as bona fide. In higher eukaryotes, GSK-3 has two isoforms, GSK-3α and

GSK-3β, and participates in a large number of cellular processes including Wnt, Notch and

Hedgehog signalling pathways, insulin regulation of glucose, neural development, cell

Page 79: Protein-kinase inhibitors as potential leishmanicidal drugs

64

division, transcription, and cell survival and death.144 GSK-3 in Leishmania spp. has a long

and a short isoform. The latter (GSK-3s) (Figure 9) is the closest homologue to the Homo

sapiens GSK-3β (HsGSK-3β), sharing 49% sequence identity. LGSK-3 is essential for the

progression of the cellular cycle, and its inhibition leads to the programmed cell death of the

parasite.111 Due to its localization in the flagellum and the basal body, it was recently

suggested that the short isoform of LGSK-3 additionally participates in the regulation of

flagellum morphogenesis and/or motility besides cell division, as it was surmised for

T. brucei GSK-3s.145 Moreover, Xingi et al.111 reported that the expression of L. donovani

GSK-3s (LdGSK-3s) in amastigotes (both axenic and isolated from spleen) is 3-fold lower

than in the promastigote stage. Since the amastigote form undergoes metabolic reprograming

and adopts a semi-quiescent state,68 the lower expression of LdGSK-3s in amastigotes may

be related to the overall downregulation of its metabolism.

Figure 9. Crystal structure of LmjGSK-3s (PDB code: 3E3P). The α-helical and β-sheet domains are

coloured in cyan and blue, respectively. Yellow colours within the active site stand for the hinge and

activation loop. The glycine-rich loop, disordered in the crystal structure, is represented with magenta-

coloured dots.140

Page 80: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

65

The GSK-3s of Leishmania became an attractive target for drug development after its

pharmacological validation by Xingi et al.111 using indirubin analogues, which are described

as potent inhibitors of mammalian CDKs and GSK-3. Among them, 5-Me-6-BIO showed

leishmanicidal activity and inhibited LdGSK-3s. This work endorsed the feasibility of

repositioning HsGSK-3β inhibitors against LGSK-3s. Since then, only two additional studies

documented a target-based approach focused on LGSK-3s; one tested other indirubin

analogues146 whereas the other evaluated commercial protein kinase inhibitors.140 Two

additional studies addressed the importance of efficient computational methods for the

virtual pre-selection of GSK-3 inhibitors for their subsequent screening in vitro.147,148

In this work, we aimed to expand the diversity of chemical scaffolds with

leishmanicidal properties and find new alternatives with a different mechanism of action

from those of current antileishmanial drugs in clinical use. For this purpose, we used small

heterocyclic molecules from diverse chemical families, some of which formerly developed

as potential human PKIs, to be screened against a recombinant GSK-3s of L. donovani.

4.1. LdGSK-3s extraction and purification

LdGSK-3s was obtained from E. coli BL21(D3) transfected with a pET28a(+) plasmid

containing the gene for LdGSK-3s. The expression vector pET28(+) is regulated by a lac

operon that includes the resistance gene to kanamycin. Insertion of the LdGSK-3s gene

within the polylinker region between Nco I and Xho I sites provides the expressed enzyme

with a C-terminal poly-(His)6 tag. The poly-His(6) tag is a standard peptide tag used for

single-step purification of recombinant proteins on immobilized metal ion affinity

chromatography, using Ni2+ or Co2+-loaded resins.149 This procedure has been used for the

purification of other Leishmania kinases besides LdGSK-3s,111 such as LmjCK1.2,118

LmxMAPK4 and LmxMKK5,150 LdAIRK1,123 LmxCRK3,151 and more recently

LdMAPK3.107 Expression and purification of recombinant LdGSK-3s was carried out

according to the work of Rachidi et al.118 with two modifications: i) disruption by French

press instead of sonication, owing to the easier control of the temperature of the process; and

ii) purification via Ni2+-column over the Co2+-column due to its slightly higher performance.

Page 81: Protein-kinase inhibitors as potential leishmanicidal drugs

66

The transfected E. coli BL21(D3) bacteria were selectively grown in 1 L of

Luria-Bertani broth with 50 μg/mL kanamycin. Expression of LdGSK-3s was induced at an

OD595nm = 0.6 by addition of 1 mM isopropyl-β-D-1-thiogalactopyranoside (IPTG)

(Figure 10, panel A). Cells were harvested by centrifugation 3 h later, and LdGSK-3s was

subsequently purified by cellular disruption of the pellet with Lysis Buffer and extrusion

through a French press, followed by solubilization of the enzyme from the lysate with 0.1%

TX-100. Afterwards, the insoluble debris from the lysate was removed by centrifugation.

The soluble LdGSK-3s was purified by affinity chromatography on a 5 mL Ni2+-column

profiting from its poly-(His)6 tag. Non-bound material and weak unspecific binding of

contaminants were removed by 15 column volumes of Washing Buffer containing 30 mM

imidazole, which competes with His residues in the binding to the Ni2+-column. Finally,

LdGSK-3s was eluted with 4 column volumes of Elution Buffer containing 300 mM

imidazole, and collected in fractions of 1.8 mL of volume. The SDS-PAGE protein patterns

of fractions and samples of each purification step are compiled in Figure 10 (panels B and

C). In these polyacrylamide gels, LdGSK-3s appears as a wide band centered at ~40 kDa.111

Page 82: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

67

Figure 10. SDS-PAGE electrophoresis of the samples from each purification step. LdGSK-3s is

highlighted in a red box. M: molecular weight marker. A) Protein pattern of E. coli

BL21(D3)-pET28a(+)LdGSK-3s before and after 3 h induction with 1mM IPTG. B) Protein pattern for

each purification step. I: cell culture (pellet) solubilized in Lysis Buffer. II: Lysate solution after French

press lysis. III: Supernatant obtained after 30 min of solubilization with 0.1% TX-100 and centrifugation

of the lysate solution. IV: Non-bound material from the supernatant. V-VIII: Samples from the

exhaustive washing of the column with Washing Buffer. C) SDS-PAGE pattern of the fractions (1-18)

eluted from the Ni2+ column with Elution buffer.

In order to remove the imidazole, the highly purified LdGSK-3s from fractions 3-18

was pooled and dialyzed in an Amicon filtration device with a 30 kDa membrane cut-off

against Kinase PBS. Next, non-soluble material was removed from the protein solution by

centrifugation. Finally, the enzyme was filtered through a 0.22 μm pore filter, and the

resulting LdGSK-3s (Figure 11; sample d, circled in green) was quantified with

bicinchoninic acid, using bovine serum albumin for the standard curve. One litre of E. coli

BL21(D3)-pET28a(+)LdGSK-3s culture (OD595nm = 0.6) after induction yielded 11.86 mg

Page 83: Protein-kinase inhibitors as potential leishmanicidal drugs

68

of LdGSK-3s. For the sake of comparison, the reported yield of recombinant LdMAPK3 was

around 2-4 mg/L.107

Figure 11. Electrophoretic pattern of the different steps of the concentration and dialysis of

LdGSK-3s. M: molecular weight marker. a: Pooled fractions (3-18) from imidazole elution. b: Dialyzed

protein retained by the Amicon, 30 kDa cut-off membrane. c: Ultracentrifuged protein (147,000 ×g, 1 h).

d: Resulting LdGSK-3s after filtration through a 0.22 μm nitrocellulose filter. e-h: Eluted samples from

the four consecutive washings in the Amicon device. i: Pellet from the ultracentrifugation step.

Next, we checked the kinase activity of the purified LdGSK-3s through a

luminescence-based activity test using the Kinase-Glo® kit (Promega), which allows the

measurement of ATP consumption by the kinase. The reagents of this kit consist of a

thermostable luciferase plus D-luciferin that emit luminescence in the presence of Mg2+, O2

and ATP (Figure 12). Addition of the Kinase-Glo® reagents to a kinase reaction quenches

the reaction and allows indirect measuring of the kinase activity through the end-point

luminescence generated by the luciferase with the remaining ATP. This method avoids less

safer alternatives, such as radioactive assays based on the incorporation of γ-32P or γ-33P into

proteins or peptide substrates, as well as more expensive and time consuming assays based

on phosphorylation-dependent changes in fluorescence or luminescence of tagged peptide

substrates and/or substrate-specific antibodies.152

Page 84: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

69

Figure 12. Mechanism of ATP measurement in the Kinase-Glo® method.153

The use of the Kinase-Glo® kit was first described for the study of HsGSK-3β

activity,153 and has been used with the GSK-3 of other organisms145 including

Leishmania,111,146 as well as for other PKs.154,155 The main limitation of this method is that

contamination with other ATP-hydrolysing enzymes leads to overestimation of the kinase

activity.152 To avoid this issue, GSK-3 specific activity was achieved using

phosphoglycogen synthase peptide 2 (GS-2, sometimes named GS-1;

YRRAAVPPSPSLSRHSSPHQ(pS)EDEEE) as substrate for LdGSK-3s (Figure 13, panel

A), as described for HsGSK-3β,156 TbGSK-3s145 and Plasmodium falciparum GSK-3.157 The

use of peptide substrates instead of proteins not only provides higher solubility, but also a

single and specific phosphorylation sequence demanded by the kinase. In fact, GSK-3

preferably recognizes and phosphorylates substrates with a prior phosphorylation under a

S/TXXXpS/T pattern,143 and the commercial GS-2 peptide fulfils this pattern, presenting a

21pS residue phosphorylated by CK-2.156

Moreover, inhibition of kinase activity with indirubin-3’-monoxime-5-sulphonic acid,

a known inhibitor of GSK-3,140 was used to confirm that the measured ATP consumption

was specific to LdGSK-3s (Figure 13, panel B). The experimental IC50 on LdGSK-3s

determined in our assays for this inhibitor (IC50 = 2.39 ± 0.22 μM) was similar to that

described on the short isoform of GSK-3 from L. infantum (LiGSK-3s,

Page 85: Protein-kinase inhibitors as potential leishmanicidal drugs

70

IC50 = 2.22 ± 0.07 μM),140 as expected from the 100% sequence identity between

L. infantum and L. donovani GSK3-s.111

Figure 13. Validation of the LdGSK3-s kinase activity. Panel A: Controls for GSK-3 kinase activity.

Panel B: IC50 for the inhibitor indirubin-3’-monoxime-5-sulphonic.

4.2. Selection of the different compounds tested

In a first step, four different sets of compounds from our in-house chemical library138

were assessed. Set 1 consisted of 27 described inhibitors of human PKs such as GSK-3β,

CK1 or LRRK2. Remarkably, besides GSK-3β, CK1 also has a corresponding counterpart

in Leishmania, named CK1.2. Set 2 to 4 encompassed compounds selected from our

chemical library after performing virtual screening on the GSK-3s,111 CK1.2118 and

MAPK4109 of Leishmania, all three validated as pharmacological targets against the parasite.

Virtual screening is a powerful tool for the selection of chemical structures with putative

binding capability to a specific given locus in the targeted protein. In Leishmania, a recent

work has already profited from this strategy to primarily identify putative inhibitors of the

Akt-like kinase of this parasite,94 while other studies used these tools to study enzyme-

inhibitor interactions in the AIRK123 and CDK12104 protein kinases of the parasite. In this

work, virtual screening was extensively used for a primary screening as well as to infer the

feasible location of the binding site inside the targeted protein. For the virtual screening on

Page 86: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

71

the GSK-3s of Leishmania, the search was focused on allosteric pockets of the enzyme using

the crystal structure of LmjGSK-3s,140 leading to Set 2 with 24 compounds. With regard to

the selection of potential inhibitors of LmjCK1.2 and LmxMAPK4, as no crystal structures

were available, homology models of both enzymes were built prior to performing a virtual

screening focused on their catalytic site. As a result, 14 compounds (Set 3) were selected as

in silico hits for the LmjCK1.2 model and 28 (Set 4) for the LmxMAPK4 model.158,159

Two additional subsets of compounds, Set 1.1 and Set 2.1, were also included in this

work, selected as a refinement of those compounds within Set 1 to 4 with the highest

performance in the initial assays for leishmanicidal and LdGSK-3s inhibition, as well as

cytotoxicity.

4.3. Assessment of the biological and inhibitory activities of the selected

compounds

The phenotypic screening and IC50 estimations for Leishmania parasites were

specifically carried out with axenic cultures of promastigotes and amastigotes in exponential

phase. Promastigotes in exponential growth phase (procyclic promastigotes) were used to

better study the effect of our compounds on the active metabolism of this proliferative stage,

in contrast from those in stationary growth phase, which are mostly constituted by non-

dividing metacyclic promastigotes with a low metabolic rate.160 The use of an established

line of axenic amastigote culture of L. (amazonensis) pifanoi ensured homogeneity and

afforded a large production of amastigotes while adhering to the 3R principle to spare animal

lives, which is not fulfilled with lesion-derived amastigotes. Axenic amastigote cultures are

obtained by induction of amastigote differentiation in axenic promastigote cultures via

stepwise increase of temperature and acidic pH besides the use of specific media. Axenic

amastigote cultures have been achieved for many Leishmania species including L. donovani,

L. tropica, L. mexicana, L. braziliensis, L. panamensis, L. amazonensis and the more

recently discovered L. orientalis.161-164 These axenic amastigotes are closely similar to

lesion-derived amastigotes, mimicking several different morphological, biological,

molecular, biochemical and immunochemical features, including their metabolic profile,

Page 87: Protein-kinase inhibitors as potential leishmanicidal drugs

72

with upregulation of stage-specific molecules and conservation of their infectivity.68,163 With

regard to their pharmacological profile, amphotericin B showed similar activity on both

axenic and lesion-derived amastigotes, whereas the axenic amastigote model lacks this

translational aspect when the activity of the drug is dependent on its activation by the

macrophage (as reported for pentavalent antimonials)165 and/or modulation of the

macrophage (as described for miltefosine).166 The amastigote axenic model was used in this

work for a primary phenotypic screening to ensure the effectivity of selected compounds

against this form of the parasite. Those hits with proliferation and LdGSK-3s inhibitory

activities were further evaluated on amastigote-infected macrophages to assess the relevance

of the host cell for their physiological leishmanicidal activity.

4.3.1. Reported inhibitors of human protein kinases (Sets 1 and 1.1)

Compounds from Set 1 (1-27) are reported inhibitors of human kinases. Compounds

1 to 11 are human GSK-3 inhibitors from different chemical families (including

thiadiazolidindione,167 iminothiadiazole,168 quinoline,169 halomethylketone,170 maleimide171

and thiazole172 derivatives) and different enzyme-binding modes. Compounds 12 to 20 are

ATP-competitive inhibitors of human CK1 bearing a benzothiazole173 or imidazole174

scaffold in their structure, and compounds 21 to 27 are indolinone-like175 LRRK2 inhibitors.

The biological activities of these compounds on Leishmania promastigotes, amastigotes, and

MPMs are compiled in Table 7.

From the phenotypic assays carried out as described in Material and Methods, only 9

out of the 27 compounds from Set 1 (compounds 1, 2, 5, 6, 8, 10, 19, 25 and 26) resulted

active on Leishmania. Their leishmanicidal activities were consistently higher on axenic

amastigotes than on promastigotes, except for the indolinone 25, with a similar activity on

both stages. For compounds 5 and 8, leishmanicidal activity was limited to the amastigote

form. Within the 9 leishmanicidal compounds, cytotoxicity on MPMs was always lower than

their respective leishmanicidal activity on amastigotes except for the benzothiazole 19.

Moreover, the highest selectivity index values (SI) were found in the thiadiazolidindione 1,

the maleimide 8 and the indolinone 26 (SIs = 16.5, >20.0 and 23.8, respectively).

Page 88: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

73

With regard to the enzymatic activity on LdGSK-3s, no inhibitory effect was observed

for the human CK1 and LRRK2 inhibitors from this set. In contrast, within the human

GSK-3 inhibitors, the thiadiazolidindiones (TDZDs) 1 and 2, the iminothiadiazoles

(ITDZs) 3 and 4, and the halomethylketones (HMKs) 6 and 7 showed inhibition of

LdGSK-3s at micro- and nanomolar range.

In order to get a deeper insight into the putative structure-activity relationship (SAR)

of active compounds from Set 1, a focused set of 20 compounds (Set 1.1) was gathered

(Table 8). This second selection included ten derivatives of the GSK-3 inhibitor 6 and

maleimide 7 (compounds 28 to 37) and ten benzothiazole derivatives of the CK1 inhibitor

19 (compounds 38 to 47). From this set, eight compounds (compounds 28-30, 32, 33, 36, 43

and 46) showed significant leishmanicidal activity (IC50s <25 μM). Only the benzothiazole

43 was more active on promastigotes than amastigotes. The best selectivity profiles were

those of compounds 28, 36 and 46 (SI = 14.6, ~11.6 and >25 respectively).

As observed for Set 1, none of the compounds from Set 1.1 that were analogues of the

CK1 inhibitor 19 had activity on the GSK-3s of Leishmania. Among the HsGSK-3

inhibitors from this set, those that act through an irreversible binding mode on the human

enzyme (compounds 28 to 33) were inhibitors of LdGSK-3s, except for compounds 30 and

31. When inhibition of LdGSK-3s was compared to the leishmanicidal activity within each

compound, only compound 33 showed a higher inhibitory effect on the enzyme than on the

parasite.

Altogether, only the TDZDs 1 and 2 and the HMK 6 from Set 1 and the HMK

derivatives 28, 29, 32 and 33 from Set 1.1 showed both leishmanicidal activity and inhibition

of LdGSK-3s. In contrast, the ITDZs 3 and 4 were potent LdGSK-3s inhibitors with IC50s at

submicromolar range, but lacked leishmanicidal activity.

Page 89: Protein-kinase inhibitors as potential leishmanicidal drugs

74 Table 7. Enzymatic and biological characterization of Set 1 (selected human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

1

1.1 ± 0.2 10.9 ± 0.4 2.0 ± 1.9 32.9 ± 3.5 16.5

2

0.32 ± 0.05 17.6 ± 2.3 7.1 ± 1.8 >50 >7.0

3

0.248 ± 0.004 >25 >50 - -

4

0.174 ± 0.001 >25 >50 - -

5

>10 >50 3.6 ± 1.3 9.9 ± 0.9 2.8

S

N

N

NNO

2HBr

Page 90: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

75

Table 7 (continued). Enzymatic and biological characterization of Set 1 (selected human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

6

1.8 ± 0.3 4.6 ± 0.2 2.2 ± 0.6 6.3 ± 1.2 2.9

7

17.7 ± 2.7 >50 >50 - -

8

>10 >50 2.5 ± 2.7 >50 >20.0

9

>10 >50 >50 - -

10

>10 >25 14.4 ± 2.6 32.0 ± 3.1 2.2

HN

O O

O

NHBr

HN

O O

O

NH

Me

MeO

NH

NH

O

S

N

Me

O

Me

MeO

NH

NH

O

S

N

Me

O

Br

Page 91: Protein-kinase inhibitors as potential leishmanicidal drugs

76 Table 7 (continued). Enzymatic and biological characterization of Set 1 (selected human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

11

>10 >50 >50 - -

12

>10 >50 >50 - -

13

>10 >50 >50 - -

14

>10 >50 >50 - -

MeO

NH

NH

O

S

N

Me

O

Br

Br

N

SNH

O

Cl

F3C

N

SNH

O

F3C

N

SNH

O

MeO

Cl

Page 92: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

77

Table 7 (continued). Enzymatic and biological characterization of Set 1 (selected human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

15

>10 >50 >50 - -

16

>10 >50 >50 - -

17

>10 >50 >50 - -

18

>10 >50 >50 - -

N

SNH

O

EtO

Cl

N

SNH

O

F3C

Cl

N

SNH

O

F3C

OMe

N

SNH

O

F3C

Cl

Cl

Page 93: Protein-kinase inhibitors as potential leishmanicidal drugs

78 Table 7 (continued). Enzymatic and biological characterization of Set 1 (selected human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

19

>10 32.8 ± 3.9 20.3 ± 1.9 >12.5 >0.6

20

>10 >25 >50 - -

21

>10 >25 >25 - -

22

>10 >25 >25 - -

N

SNH

O

Cl

Cl

N

N

Page 94: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

79

Table 7 (continued). Enzymatic and biological characterization of Set 1 (selected human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

23

>10 >50 >50 - -

24

>10 >50 >50 - -

25

>10 4.1 ± 0.1 4.5 ± 1.5 >25 >5.6

26

>10 22.9 ± 2.6 2.1 ± 0.2 ~50* ~23.8

27

>10 >50 >50 - -

* The viability of macrophages was 55.2 ± 6.1 % for compound 26 at 50 μM

NH

N

O

NH

Me

Page 95: Protein-kinase inhibitors as potential leishmanicidal drugs

80 Table 8. Enzymatic and biological characterization of Set 1.1 (second selection of human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

28

7.5 ± 1.4 6.6 ± 0.5 0.5 ± 0.1 7.3 ± 1.0 14.6

29

10.3 ± 2.0 4.4 ± 0.1 2.4 ± 0.4 3.6 ± 0.5 1.5

30

>10 14.2 ± 1.0 1.2 ± 0.2 3.2 ± 0.4 2.7

31

>10 >50 >50 - -

32

3.0 ± 0.4 8.6 ± 0.3 1.2 ± 0.2 7.9 ± 0.6 6.6

33

1.6 ± 0.2 >50 6.5 ± 2.0 >25 >3.8

HN

O O

O

N

MeBr

Page 96: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

81

Table 8 (continued). Enzymatic and biological characterization of Set 1.1 (second selection of human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

34

>10 >50 >25 - -

35

>10 >50 29.5 ± 9.8 >25 >0.8

36

>10 >50 4.3 ± 0.2 ~50* ~11.6

37

>10 >50 >50 - -

* The viability of macrophages was 51.0 ± 1.9 % for compound 36 at 50 μM.

HN

O O

O

NMe

Me

HN

O O

O

NH

NMe

HN

O O

O

NNMe

Me

HN

O O

O

NH

Me Me

Page 97: Protein-kinase inhibitors as potential leishmanicidal drugs

82 Table 8 (continued). Enzymatic and biological characterization of Set 1.1 (second selection of human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

38

>10 >50 >50 - -

39

>10 >50 >50 - -

40

>10 >50 >50 - -

41

>10 >50 >50 - -

N

SNH

O

MeO

OMe

N

SNH

O

EtO

OMe

N

SNH

O

F3C

OMe

N

SNH

O

F3C

OMe

Page 98: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

83

Table 8 (continued). Enzymatic and biological characterization of Set 1.1 (second selection of human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

42

>10 >50 >50 - -

43

>10 8.1 ± 1.6 13.3 ± 1.7 4.1 ± 2.8 0.3

44

>10 >50 >50 - -

N

SNH

NHO

F3C

OMe

N

SNH

NHO

F3C

Cl

N

SNH

NHO

Me

OMe

Me

O

Page 99: Protein-kinase inhibitors as potential leishmanicidal drugs

84 Table 8 (continued). Enzymatic and biological characterization of Set 1.1 (second selection of human protein kinase inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

45

>10 >50 >50 - -

46

>10 6.5 ± 0.7 1.0 ± 0.1 >25 >25.0

47

>10 >50 >50 - -

N

SNH

F3C

OMe

N

SNH

O

Me

EtO

O

N

S

Me

NH

EtO

O

OCl

Page 100: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

85

Additional computational studies were performed to decipher the binding mode of the

LdGSK-3s inhibitors from these sets and explain the lack of activity of some HsGSK-3β

inhibitors. TDZDs 1 and 2 are well-known non-ATP competitive HsGSK-3β inhibitors167

that bind covalently to the thiol group of Cys199 at the active site of GSK-3,176 as also

described for the HMK family171 (6, 28, 29, 32, as well as compound 33, a maleimide with

a HMK unit). Since the Cys199 is conserved in the Leishmania GSK-3s as Cys169, this

residue was expected to behave as the acceptor group for the covalent inhibition of

LdGSK-3s by compounds 1, 2, 6, 28, 29, 32 and 33. In contrast, the ITDZs 3 and 4 are

described substrate-competitive HsGSK-3β inhibitors.168 As the binding region for the

substrate in the GSK-3 in Leishmania is rather similar to the human one, compounds 3 and

4 were surmised to interact with LdGSK-3s through this site. These hypotheses were

subsequently confirmed by different molecular docking studies carried out on the crystal

structure of LmjGSK-3s140 with compounds 2, 3, 4 and 6 as representative molecules (Figure

14). The maleimide 34, an HsGSK-3β reversible inhibitor with nil inhibitory activity on

LdGSK-3s, was also included in the molecular docking studies for the sake of comparison

with the maleimide 33, an HsGSK-3β irreversible inhibitor with LdGSK-3s inhibition. Both

maleimides are reported as ATP-competitive HsGSK-3β inhibitors with almost identical

structure, differing in the presence of the HMK tail in compound 33 with respect to its

absence in compound 34.171 It was identified that the lack of LdGSK-3s inhibition by

maleimide 34 obeys the restricted access of this compound into the ATP binding cavity due

to the Met100 residue of LGSK-3s, which is a bulkier gatekeeper than Leu132 in HsGSK-3β

(Figure 15). Additionally, docking studies with quinoline 5, an HsGSK-3β allosteric

inhibitor, showed that its inability to inhibit LdGSK-3s was caused by the replacement of

the flexible and cationic Arg209 from the human kinase with the neutral and rigid Pro178 in

the leishmanial GSK-3s.177

Page 101: Protein-kinase inhibitors as potential leishmanicidal drugs

86

Figure 14. Binding mode into the LmjGSK-3 enzyme for ITDZ, TDZD, and HMK representative

compounds.177 A) Blind docking poses obtained from the most representative clusters for 2 (TDZD) and

4 (ITDZ) in the ATP binding site and the substrate binding site, respectively. B) Superimposition of most

representatives regular docking results of ITDZ compounds 3 and 4. C) Superimposition of the best

covalent docking poses obtained for TDZDs (1 and 2) and HMK 6. D) Detailed view of the covalent

docking for compound 2.

Page 102: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

87

Figure 15. A) Superposition of the human (purple) and Leishmania (green) GSK-3 enzymes. Key

mutations in the ATP binding site are depicted as sticks. B) Superposition of the binding poses of

maleimide 34 to the human (purple) and Leishmania (green) GSK-3 enzymes.177

The reactivity of covalent inhibitors like TDZDs 1 and 2 as well as the HMKs 6, 28,

29, 32 and 33 towards thiol groups may cause cross-inhibition with thiol-dependent

enzymes. In Leishmania, such enzymes include the dimeric trypanothione reductase,

tryparedoxin and tryparedoxin peroxidase from the thiol-based redox system,80 adenosine

kinase from the purine savage pathway,178 carbonic anhydrases179 and cysteine proteases

among others. Remarkably, antimonials also have high affinity for thiol containing

biomolecules,180 suggesting a feasible synergy with GSK-3 inhibitors. As a downside, owing

to their potential unspecific-reactivity with nucleophilic residues, covalent inhibitors are

commonly defined as PAINs (pan-assay interference compounds). PAINs usually appear as

very promising hits by interfering in the readout of many biological assays,181 thus they are

usually removed from initial screenings. Nevertheless, the number of drugs acting through

covalent binding is far from scarce, and includes FDA approved drugs such as salicylic acid,

β-lactam antibiotics and several antitumoral drugs.182 A crucial aspect for the effectiveness

and safety of irreversible inhibitors is the turnover rate of their targets. This is especially

relevant in infectious diseases where the same compound can target both the pathogen

enzyme and its human counterpart, as proven by eflornithine (α-difluoromethylornithine;

Page 103: Protein-kinase inhibitors as potential leishmanicidal drugs

88

DFMO). DFMO, an anti-trypanosomatid drug for African trypanosomiasis, is an irreversible

inhibitor of the ornithine decarboxylase (ODC) enzyme, a key regulator of the polyamine

biosynthesis pathway. DFMO has similar affinity and effective inhibition on both the human

and the T. brucei gambiense ODC, but the slower turnover rate (t1/2 over 6 h) of the parasitic

enzyme with respect to its human homolog (t1/2 <1 h) endows DFMO with inhibitory

specificity for the ODC of T. brucei.183 The turnover rate of the GSK-3s in Leishmania is

still unknown, whereas the reported half-life of GSK-3β in murine neurons is 41 (± 4)176 -

48184 h. Since intracellular amastigotes have higher generation time (33 h),185 achieving

selective effect on the GSK-3s of Leishmania amastigotes due to turnover rate differences

seems unlikely.

Nevertheless, when the in vitro inhibition of LdGSK-3s is compared with that of

HsGSK-3β (Table 9), the TDZDs 1 and 2, and the ITDZs 3 and 4 had higher inhibitory

activity on LdGSK-3s than on the human one, thus being appealing candidates for further

optimization.

Table 9. Inhibition of HsGSK-3β of the leishmanicidal LdGSK-3s inhibitors from Sets 1 and 1.1

(selections of human protein kinase inhibitors).

Compound HsGSK-3β

IC50 (μM)

LdGSK-3s

IC50 (μM)

1 2167 1.1 ± 0.2

2 0.7 ± 0.1186 0.32 ± 0.05

3 0.88168 0.248 ± 0.004

4 1.95168 0.174 ± 0.001

6 0.5170 1.8 ± 0.3

28 1.0170 7.5 ±1.4

29 5.0170 10.3 ±2.0

32 1.0170 3.0 ± 0.4

33 0.005 ± 0.001171 1.6 ± 0.2

Page 104: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

89

Furthermore, the TDZD 2 is also known as tideglusib, and is currently in clinical trials

as an HsGSK-3β inhibitor for neurological disorders (phase II), with little to no signs of

toxicity at daily doses ranging from 400 to 1,000 mg.187 Hence, owing to its toxicity profile

and higher activity on LdGSK-3s, tideglusib is an appealing candidate from drug

repurposing.

4.3.2. In silico inhibitors of LmjGSK-3s (Set 2), LmjCK1.2 (Set 3) and

LmxMAPK4 (Set 4)

As aforementioned, Xingi et al.111 first demonstrated the successful use of HsGSK-3β

inhibitors as a source for Leishmania GSK-3s inhibitors when they reported the finding of

LdGSK-3s inhibitors among some 6-bromo indirubins, which are strong inhibitors of

mammalian CDKs and GSK-3. While none of the compounds had higher affinity for

LdGSK-3s over the mammalian GSK-3β, Xingi and co-authors suggested focusing on amino

acid differences at the binding site of the enzyme as an approach for hit optimization. Ojo

and co-authors140 also demonstrated the appeal of drug repurposing by demonstrating

cross-inhibition of commercial HsGSK-3β inhibitors on the GSK-3s of different Leishmania

species. Only TDZDs 1 and 2 and ITDZs 3 and 4 from our screening of Set 1 had higher

activity on the Leishmania GSK-3s over HsGSK-3β (Table 9), thus we were prompted to

carry out virtual screenings of our chemical library (constituted by more than 2,000

compounds) for the selection of new groups of compounds. Virtual screening is a useful

technique to carry out massive target-based screenings of compounds on a targeted protein

when its structure, or at least the binding sites, are well defined.188 In Leishmania, this

approach was endorsed by the finding of two specific inhibitors in vitro of the Akt-like

kinase of L. panamensis, primarily selected from the virtual screening of a 600,000-

compound library focused on the regulatory pleckstrin domain of this modelled kinase.94

Thus, Sets 2 to 4 were selected from our in-house chemical library by virtual

screenings. To this end, the targets were expanded from the initial GSK-3s, to CK1.2 and

MAPK4, two additional pharmacologically validated protein kinases of Leishmania.109,118

Page 105: Protein-kinase inhibitors as potential leishmanicidal drugs

90

Set 2 (compounds 48-71) was originated from the virtual screening against the

allosteric pockets of LmjGSK-3s,159 using the crystal structure of the enzyme provided by

Ojo et al.140 The lower conservation of allosteric pockets among protein kinases with their

active site opens a broader avenue for the finding of specific inhibitors. In Leishmania, the

successful exploration of allosteric pockets was reported through the identification of 4,6-

diamino substituted pyrazolopyrimidines as new inhibitors against L. major

methionyl-tRNA synthetase (MetRS), a key enzyme in protein synthesis.189

Our virtual screening on the allosteric pockets of LmjGSK-3s identified 24 compounds

with a variety of chemical scaffolds, and their biological activities were subsequently

evaluated (Table 10). From this set, only the quinone 48 showed a relatively significant

leishmanicidal activity on both stages besides LdGSK-3s inhibition (IC50s ca 10 µM for all

three activities).

Quinones are typical inhibitors of the respiratory chain, antagonizing the function of

mitochondrial ubiquinone and promoting oxidative stress.190,191 They are also considered

PAINs due to their redox-cycling properties and reactivity with nucleophilic residues.

Despite this, quinones such as the antineoplastic anthracyclines and the antitrypanosomatids

atavoquone and buparvaquone, among others, are nowadays in clinical use.181,191 Moreover,

naphthoquinones reportedly exhibit a wide variety of effects, including anticancer, antiviral,

immunomodulatory, trypanocidal and antimicrobial activities, due to inhibition of

respiration or covalent interaction with their targets.192

Hence, we were prompted to synthesize new naphthoquinone compounds to further

explore and optimize this scaffold.159 The biological activities of 23 new naphthoquinone

derivatives 72-94 (Set 2.1) are compiled in Table 11.

Page 106: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

91

Table 10. Enzymatic and biological characterization of Set 2 (in silico LmjGSK-3s inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

48

~10* 10.5 ± 1.2 11.2 ± 2.5 ND ND

49

>10 >50 >50 - -

50

>10 >50 >50 - -

51

>10 >50 >50 - -

* Compound 48 at 10 μM causes 45.7 ± 7.6 % inhibition of LdGSK-3s activity.

N

SNH

OO O

NH

Me

O

NH

O

HNO

N

HN

O

NHO

Page 107: Protein-kinase inhibitors as potential leishmanicidal drugs

92 Table 10 (continued). Enzymatic and biological characterization of Set 2 (in silico LmjGSK-3s inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

52

>10 >25 >25 - -

53

>10 >50 >50 - -

54

>10 >50 >50 - -

55

>10 >50 >50 - -

56

>10 >50 >50 - -

NH

NH

S N

N CF3

Page 108: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

93

Table 10 (continued). Enzymatic and biological characterization of Set 2 (in silico LmjGSK-3s inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

57

>10 >50 >50 - -

58

>10 >50 >50 - -

59

>10 >50 >50 - -

60

>10 >50 >50 - -

61

>10 >25 >50 - -

N

N

O

S

Et

F

F

Cl

Page 109: Protein-kinase inhibitors as potential leishmanicidal drugs

94 Table 10 (continued). Enzymatic and biological characterization of Set 2 (in silico LmjGSK-3s inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

62

>10 >50 >50 - -

63

>10 >50 >50 - -

64

>10 >50 >50 - -

65

>10 >50 20.8 ± 0.0 >25 >1.2

66

>10 >50 >50 - -

N

SNH

OO O

NH

OPh

OH

N

Me

O

NH

OHN

O6

OH

N

Me

O

NH

OHN

O

O

Ph

O

Page 110: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

95

Table 10 (continued). Enzymatic and biological characterization of Set 2 (in silico LmjGSK-3s inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

67

>10 >50 >50 - -

68

>10 >50 >50 - -

69

>10 >50 >50 - -

70

>10 >50 >50 - -

71

>10 >50 >50 - -

N

S

Me

NH

EtO

O

O

OMe

OMe

OMe

NH

NH

O

NOMe

OH

N

Me

O

NH

OHN

O

N

Page 111: Protein-kinase inhibitors as potential leishmanicidal drugs

96 Table 11. Enzymatic and biological characterization of Set 2.1 (new quinone analogues).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

72

2.5 ± 0.1 1.51 ± 0.00 0.51 ± 0.00 2.6 ± 0.0 5.1

73

>10 1.97 ± 0.20 0.4 ± 0.0 >10 >25

74

>10 >25 10.6 ± 1.0 4.73 ± 0.60 0.4

75

3.7 ± 0.3 1.54 ± 0.00 0.15 ± 0.40 3.1 ± 0.1 20.6

76

2.5 ± 0.2 1.47 ± 0.00 0.4 ± 0.2 4.51 ± 0.30 11.2

Page 112: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

97

Table 11 (continued). Enzymatic and biological characterization of Set 2.1 (new quinone analogues).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

77

4.5 ± 0.6 3.8 ± 1.9 3.6 ± 0.6 8.1 ± 1.0 2.3

78

4.1 ± 0.1 2.9 ± 0.2 3.7 ± 1.1 7.4 ± 0.8 2.0

79

5.3 ± 1.0 5.3 ± 0.6 1.2 ± 0.4 13.0 ± 2.8 10.8

80

2.8 ± 0.1 4.9 ± 0.2 4.7 ± 0.8 9.8 ± 0.2 2.1

81

>10 19.8 ± 3.6 17.8 ± 1.1 17.0 ± 0.3 1.0

Page 113: Protein-kinase inhibitors as potential leishmanicidal drugs

98 Table 11 (continued). Enzymatic and biological characterization of Set 2.1 (new quinone analogues).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

82

>10 1.4 ± 0.4 2.1 ± 0.6 1.3 ± 0.3 0.6

83

>10 16.9 ± 0.4 >25 - -

84

>10 19.1 ± 1.2 >25 - -

85

>10 >25 >25 - -

86

>10 >50 >50 - -

Page 114: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

99

Table 11 (continued). Enzymatic and biological characterization of Set 2.1 (new quinone analogues).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

87

>10 >25 >25 - -

88

>10 4.5 ± 0.5 1.2 ± 0.1 4.0 ± 0.2 3.3

89

>10 5.9 ± 1.8 1.8 ± 0.3 4.3 ± 0.3 2.4

90

>10 7.1 ± 2.1 3.2 ± 0.1 7.5 ± 0.6 2.3

Page 115: Protein-kinase inhibitors as potential leishmanicidal drugs

100

Table 11 (continued). Enzymatic and biological characterization of Set 2.1 (new quinone analogues).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

91

>10 >50 >50 - -

92

>10 >25 >25 - -

93

>10 4.1 ± 0.8 1.7 ± 0.2 3.8 ± 0.3 2.2

94

>10 4.5 ± 0.7 1.7 ± 0.1 4.2 ± 0.2 2.5

Page 116: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

101

Eighteen compounds (72-84, 88-90, 93 and 94) showed leishmanicidal activity in one

or both stages of the parasite. The trend towards a higher and preferential leishmanicidal

activity on the amastigote from previous sets was not fully maintained; quinones 77, 78, 80-

82 showed rather similar values for both stages, and activity on amastigotes was not

significant for compounds 83 and 84. The selectivity indexes were higher than 10 for

compounds 73, 75, 76 and 79 (SI= >25, 20.6, 11.2 and 10.8 respectively). Noteworthy,

quinones 72, 75-80 behaved as LdGSK-3s inhibitors with IC50s at low micromolar values.

While compound 48 (Set 2), a 6,7-dichloronapthoquinone, had IC50s around 10 μM

for LdGSK-3s inhibition and antileishmanial activity in both stages, its derivative

2-chloro-3-methoxy-1,4-naphthoquinone (compound 73, Set 2.1) achieved higher

leishmanicidal activities (IC50s = 1.97 ± 0.20 μM for promastigotes, 0.4 ± 0.0 μM for

amastigotes) and SI (>25), but was devoid of inhibitory activity on LdGSK-3s. From the

remaining 22 2-amino-1,4-naphthoquinones from Set 2.1, a trend for the substitution pattern

was observed. Quinones with an amine group (74, 81, 83-87, 91-92) were mostly inactive

against the parasite, whereas those with a urea (82), an amide (88-90), or a carbamate moiety

(72, 75-80, 93, 94) showed high leishmanicidal activities on both Leishmania stages. While

the presence of a chlorine atom at position 3 was not essential for leishmanicidal activity (72

vs 93 and 75 vs 94), its combination with a carbamate moiety (compounds 72, 75-80) was

strongly associated with LdGSK-3s inhibition. The LdGSK-3s inhibitors 72, 75, 76 and 80

did not show HsGSK-3β inhibition at 10 μM (19%, 30%, 25% and 19% of kinase inhibition,

respectively), with quinone 75 showing the highest SI (= 20.6). These findings were of

utmost importance because, to our knowledge, they are the first LdGSK-3 inhibitors with

scarce affinity for HsGSK-3β described so far.

Finally, Sets 3 (compounds 95-108) and 4 (compounds 109-136) were obtained from

virtual screenings on the catalytic site of homology models of LmjCK1.2 and LmxMAPK4

respectively.158,159

The CK1.2 of Leishmania was of special interest since inhibition of its human

counterpart, CK1δ, was reported for some compounds from our chemical library.173 The

search for feasible inhibitors of the Leishmania CK1.2 was exclusively guided by virtual

screening studies due to unsuccessful attempts to obtain a recombinant LmjCK1.2 in a

Page 117: Protein-kinase inhibitors as potential leishmanicidal drugs

102

soluble and active form. The lack of crystal structures for LmjCK1.2 was sorted out by

building a homology model of the enzyme. The virtual screening on the ATP binding site of

the modelled LmjCK1.2 led to the identification of 14 potential inhibitors from different

chemical families (Set 3)158 that were subsequently assayed for their biological and

inhibitory activities (Table 12).

The MAPK4 is a genetically validated target against Leishmania that plays a vital role

in stage-differentiation. Thus, a virtual screening focused on the active site of an

LmxMAPK4 model was carried out, aiming to broaden the potentiality of our in-house

chemical library. Virtual screenings on modelled Leishmania MAPK4s have been reported

before, but no additional biological studies endorsed the theoretical results.106,193 Set 4 was

formed with 28 compounds from different chemical families identified in silico as

LmxMAPK4 inhibitors. Results from the subsequent enzymatic and biological assays are

compiled in Table 13.

Unsurprisingly, none of the compounds from Set 3 or 4 inhibited LdGSK-3s, although

a cross-inhibition with other PKs is especially feasible when the highly preserved ATP

binding site is targeted.194

Only the pteridine 100 out of the 14 compounds from Set 3 was significantly active

on the parasite, with an IC50 near 6.5 μM for both amastigote and promastigote stages and a

selectivity index of 8.1. As aforementioned, Leishmania salvages pteridines from its hosts

for its folate metabolism.78 Thus, pteridine 100 may interfere with folate biosynthesis by

inhibition of the pteridine reductase PTR1 of the parasite. PTR1 is a key enzyme in the

pteridine salvaging pathway that also substitutes the dihydrofolate reductase-thymidylate

synthase (DHFR-TS) in folate reduction when DHFR-TS is inhibited.195 As such, pteridine

100 may additionally inhibit PTR1 aside from its presumed inhibition of LmjCK1.2, and

may be synergetic with DHFR-TS inhibitors against the parasite.

Page 118: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

103

From Set 4, six compounds (112, 120, 121, 124, 131 and 133) showed leishmanicidal

activity. Only the benzothiophene 121 was more active on promastigotes than on

amastigotes, whereas compounds 112, 124 and 133 had no effect on the promastigote stage.

The thiazole 133 showed the highest SI (>11.5) of the set.

Page 119: Protein-kinase inhibitors as potential leishmanicidal drugs

104

Table 12. Enzymatic and biological characterization of Set 3 (in silico LmjCK1.2 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

95

>10 >50 25.8 ± 5.9 >25 >0.6

96

>10 >50 >50 - -

97

>10 >50 >50 - -

98

>10 >50 25.1 ± 2.9 >25 >1.0

99

>10 >50 >50 - -

NH

N

O

O OMe

Page 120: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

105

Table 12 (continued). Enzymatic and biological characterization of Set 3 (in silico LmjCK1.2 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

100

>10 6.6 ± 0.3 6.2 ± 0.0 >50 >8.1

101

>10 >50 >50 - -

102

>10 >50 >50 - -

103

>10 >50 >50 - -

104

>10 >50 >50 - -

OMe

MeO OMe

NH

O

O

O2N

Me

N

N

O

S

NH2

Page 121: Protein-kinase inhibitors as potential leishmanicidal drugs

106

Table 12 (continued). Enzymatic and biological characterization of Set 3 (in silico LmjCK1.2 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

105

>10 >50 >25 - -

106

>10 >50 >50 - -

107

>10 >50 >50 - -

108

>10 >50 >50 - -

NH

NH

HN

ON

N

Me

O

N

N

N

O

S

N

Cl

Page 122: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

107

Table 13. Enzymatic and biological characterization of Set 4 (in silico LmxMAPK4 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

109

>10 >50 >50 - -

110

>10 >50 >25 - -

111

>10 >50 >50 - -

112

>10 >50 15.2 ± 1.6 >25 >1.6

113

>10 >50 >50 - -

NH

NH

S

N

NH

O

S

NMe

Ph

MeO

NH

O

N

N

Cl

Page 123: Protein-kinase inhibitors as potential leishmanicidal drugs

108

Table 13 (continued). Enzymatic and biological characterization of Set 4 (in silico LmxMAPK4 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

114

>10 >50 >50 - -

115

>10 >50 >50 - -

116

>10 >50 >50 - -

117

>10 >50 >50 - -

118

>10 >50 >25 - -

NNH

N

O

S

Page 124: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

109

Table 13 (continued). Enzymatic and biological characterization of Set 4 (in silico LmxMAPK4 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

119

>10 >50 >50 - -

120

>10 11.0 ± 0.7 3.0 ± 0.3 14.3 ± 0.9 4.8

121

>10 3.7 ± 0.4 11.8 ± 0.7 >25 >2.1

122

>10 >50 >50 - -

NH

O

Cl

N

N

Page 125: Protein-kinase inhibitors as potential leishmanicidal drugs

110

Table 13 (continued). Enzymatic and biological characterization of Set 4 (in silico LmxMAPK4 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

123

>10 >50 >50 - -

124

>10 >50 13.1 ± 4.3 >50 >3.8

125

>10 >50 >50 - -

126

>10 >50 >50 - -

N

SNH

HNO

Me

EtO

OO

Page 126: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

111

Table 13 (continued). Enzymatic and biological characterization of Set 4 (in silico LmxMAPK4 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

127

>10 >50 >50 - -

128

>10 >50 >50 - -

129

>10 >50 >50 - -

130

>10 >50 >50 - -

OMe

MeO OMe

O

O

O

O2N

Cl

OMe

MeO OMe

O

O

O

Cl

Cl

OMe

MeO OMe

O

O

O

Cl

OH

N

Et

O

NH

OHN

O

OH

Page 127: Protein-kinase inhibitors as potential leishmanicidal drugs

112

Table 13 (continued). Enzymatic and biological characterization of Set 4 (in silico LmxMAPK4 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

131

>10 9.1 ± 0.5 3.8 ± 0.6 8.8 ± 1.5 2.3

132

>10 >50 >50 - -

133

>10 >50 1.3 ± 0.1 >15 >11.5

134

>10 >50 >50 - -

OMe

MeO OMe

O

O

O

Cl

CF3

Page 128: Protein-kinase inhibitors as potential leishmanicidal drugs

Resu

lts and d

iscussio

n

113

Table 13 (continued). Enzymatic and biological characterization of Set 4 (in silico LmxMAPK4 inhibitors).

Compound Chemical formula LdGSK-3s

IC50 (μM)

L. infantum

promastigotes

IC50 (μM)

L. pifanoi

axenic

amastigotes

IC50 (μM)

Murine

peritoneal

macrophages

IC50 (μM)

Selectivity

index (SI)

135

>10 >50 >50 - -

136

>10 >50 >50 - -

OMe

MeO OMe

O

O

O NO2

Page 129: Protein-kinase inhibitors as potential leishmanicidal drugs

114

4.3.3. Leishmanicidal activity of LdGSK-3s inhibitors on intracellular

amastigotes

Aiming to assess their pharmacological value in a scenario closer to a real infection,

the 14 compounds with inhibitory activity on LdGSK-3s and leishmanicidal activity on

axenic amastigotes (TDZDs 1 and 2, HMKs 6, 28, 29 and 32, maleimide with HMK tail 33,

and quinones 72, 75-80; Tables 7, 8 and 11) were subsequently assayed on intracellular

parasites.

The activity of a given drug on the intracellular amastigote with respect to the axenic

model may differ not only due to traffic restrictions to reach the parasitophorous vacuole,

but also due to the physiology and response of the macrophage as a host cell. These effects

are of especial relevance for GSK-3 inhibitors, as cross-inhibition between the parasite and

mammalian homologs is observed for most of these inhibitors, as referred to in Sections

4.3.1 and 4.3.2.

The chosen intracellular model obeys two factors; firstly, the use of BALB/c MPMs

instead of a cellular monocytic line avoids an overestimation of the cytotoxicity. Some

human GSK-3 inhibitors have been reported as antitumoral agents.196,197 Thus, their toxicity

on tumoral monocytic cell lines is foreseeable, but avoided with primary cells. There are

reports on variation of drug effectiveness, initial infection levels, and infection development

according to the source of macrophages, including murine bone marrow macrophages and

human peripheral blood monocytes induced to macrophages198,199 besides MPMs. In our

case, MPMs were chosen for their macrophage yield, their quick parasite uptake and their

optimal support for amastigote survival and replication.200 Secondly, axenic amastigotes

from L. pifanoi (L. amazonensis) species were used due to their corroborated similarity to

those obtained from infected macrophages or animal lesions,163 affording a direct

comparison of a given drug on both amastigote types. In addition, the use of a recombinant

strain expressing the fluorescent protein mCherry facilitated quantification of the infection,

as shown in Figure 16 with representative images of infections treated with compound 6.

Page 130: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

115

Figure 16. Leishmanicidal activity of compound 6 on intracellular L. pifanoi amastigotes.

Representative fields of BALB/c murine peritoneal macrophages infected with mCherry-L. pifanoi axenic

amastigotes. Panels A and B: Untreated macrophages. Panels C and D: Infected macrophages treated with

compound 6 at 4 μM (near 2-fold its IC50 for axenic amastigotes). Panels A and C: Phase microscopy.

Panels B and D: Fluorescence microscopy (540-580 nm excitation filter). Red arrows: intracellular

amastigotes. Magnification bar: 20 μm.

Infected macrophages were treated with the compounds at an equipotent concentration

within their IC50 for amastigotes and their IC50 for MPM for 16-20 h. Statistical significance

was calculated by p values obtained with Student’s t-test.

The HMK 6, its analogue compound 32, and the quinones 72, 76 and 79 caused a

significant decrease in the parasite load beyond their respective IC50 with respect to axenic

amastigotes (near 2-fold their IC50 for compounds 6 and 72, 5-fold for compounds 32 and

76, and 10-fold for compound 79) (Figure 17).

Page 131: Protein-kinase inhibitors as potential leishmanicidal drugs

116

Quinone 76 at 2 μM induced the highest decrease (76.3 %), followed by HMK 32

(65.7 % at 6 μM) and HMK 6 (63.3 % at 4 μM). Quinones 72 at 1 μM and 79 at 12 μM

caused a decrease of 39.1 and 37.9 % respectively.

Figure 17. Percentage variation of the parasite load of BALB/c murine peritoneal macrophages

infected with mCherry-L. pifanoi axenic amastigotes after treatment with the selected LdGSK-3s

inhibitors. Student’s t-test (*: p <0.05, **: p <0.01, ***: p <0.001).

Among these 5 compounds with statistically significant effects on intracellular

amastigotes, quinone 76 and HMK 32 had the highest effects and the best selectivity profiles

(SI = 11.2 and 6.6, respectively). Thus, these two compounds will be the best candidates for

further hit optimization.

The decrease and loss of effectiveness in intracellular infections for most of the

leishmanicidal compounds is a typical example of the high attrition experimented in the

pipeline for drug development in Leishmania. It is also a demonstration of how, as

Page 132: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

117

aforementioned, the macrophage takes part in the final outcome of a chemotherapeutic

treatment, not only by restrictions in the access of the drug to the parasitophorous vacuole,

but also by possible metabolic modifications of the drug, or through unexpected side effects

on its own physiology, already hijacked by the parasite.201

Nonetheless, leishmanicidal compounds with poor or nil in vivo activity can be

considered for nanoparticle encapsulation. Nanoencapsulation consists of the vehiculation

of drugs in particles of 10-1,000 nm of either organic or synthetic nature, such as liposomes,

polymer nanoparticles, solid lipid nanoparticles, and nanosuspensions. The endocytic system

of the macrophage captures the nanoparticles and drives their accumulation into the

parasitophorous vacuole, where Leishmania dwells. Besides increasing drug concentration

in the parasitophorous vacuole and decreasing off-side effects in the host cell,

nanoencapsulation also facilitates a preferential accumulation into the macrophages

reducing toxicity issues, and protects the drug from biological degradation. A variety of

vehicles and cargo leishmanicidal molecules has been used under this approach, including

liposome nanoencapsulation for meglumine antimoniate, and polymeric and solid lipid

nanoparticles for paromomycin and AmB. The liposomal formulation of AmB (Ambisome)

is probably the most well-known.202,203

4.4. Off-target effects of protein kinase inhibitors

Although target-based approaches are driven by the “one target, one drug” motto,

off-target effects from cross-inhibition are especially feasible for PK inhibitors, since many

of them bind to the active site of their targeted kinase, and the active site is highly conserved

among protein kinases.194 Indeed, some FDA-approved PKIs such as imatinib, sorafenib and

sunitinib are multi-kinase inhibitors.204 Compounds with a reactive group for nucleophilic

acceptors, such as TDZDs and HMKs, likely find off-targets on thiol-dependent

enzymes.80,178,179 Similarly, due to their chemical reactivity, quinones are also strong

candidates for a multi-target leishmanicidal mechanism involving the respiratory chain, as

reported for naphthoquinone derivatives in T. brucei.205

Page 133: Protein-kinase inhibitors as potential leishmanicidal drugs

118

Despite increasing uncertainty about its mechanism of action, unexpected multi-target

effects may be essential for the effectiveness of a compound. Since effectiveness is the one

essential trait for drug development, multi-target compounds can be advantageous,

especially as they avoid resistance by target mutation.206,207 An example of a multi-target

drug in Leishmania is miltefosine, which primarily targets phospholipid metabolism, but

also causes direct inhibition of cytochrome c oxidase and Akt-like kinase.208 Although

miltefosine resistance is easily induced in vitro and reported in clinical settings, the

mechanisms identified describe a faulty accumulation of the drug, either by dysfunction of

its unique transporter (the aminophospholipid translocase Ldmt) or by overexpression of

efflux pumps, but not by target mutation.25

As drug resistances in Leishmania keep emerging, finding multi-target drug candidates

is a very desirable outcome that may increase the likelihood of avoiding some mutation-

induced resistances. Thus, our next objective was to explore feasible off-targets for our

leishmanicidal GSK-3 inhibitors.

4.4.1. Energy metabolism of Leishmania as an off-target effect of PKIs

The energy metabolism of the parasite is an excellent parameter to gauge its

intracellular homeostasis. In Leishmania, oxidative phosphorylation is the main source of

ATP synthesis, thus its inhibition leads to an irreversible bioenergetic collapse of the

parasite.63 In mammalian cells, GSK-3β is known to play a role in energy metabolism by

regulating the expression of metabolic proteins, the intermediary metabolism, and

mitochondrial function.209 GSK-3β is involved in the downregulation of mitochondrial

energy production through the modulation of the respiratory chain and the inhibition of

pyruvate dehydrogenase, among other substrates.210 Moreover, GSK-3β tips the balance

between the anabolic and catabolic metabolisms owing to its inhibitory activity on AMPK

(AMP-activated protein kinase), a crucial PK in the regulation of cellular energetics.

Hence, the effects of our leishmanicidal LdGSK-3s inhibitors on the energy

metabolism of Leishmania were evaluated. In order to prevent misleading conclusions and

Page 134: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

119

avoid variations in the energy metabolism caused as a secondary effect of the compounds,

assays were carried out at short incubation times.

4.4.1.1. Variation of the intracellular levels of ATP in L. donovani promastigotes by the

leishmanicidal LdGSK-3s inhibitors

The intracellular levels of ATP are the main parameter of energy metabolism of almost

any cell, including Leishmania parasites. Additionally, pyrophosphate is the main energy

coin across the exclusive acidocalcisome metabolism of Trypanosomatidae.211

In order to discriminate direct from indirect inhibition of ATP synthesis (e.g.

reprogramming of gene expression triggered by inhibition of GSK-3s, or other off-target

effects), variations in the intracellular ATP content of living parasites were monitored in

real-time after compound addition.

To this end, promastigotes of the 3-LUC L. donovani strain were used.139 This strain

is transfected with a cytoplasmic form of luciferase, which endows the parasite with in vivo

emission of luminescence in the presence of D-luciferin and ATP. The poor permeability of

D-luciferin across the plasma membrane at physiological pH was overcome by the use of a

caged luciferin analogue, DMNPE D-luciferin, making the cytoplasmic ATP the limiting

substrate of luminescence (Figure 18). Untreated parasites were used to monitor the intrinsic

decrease of luminescence during the assays. A decrease in the intracellular ATP levels can

be due to either inhibition of mitochondrial biosynthesis or permeabilization of the plasma

membrane by the compounds. Thus, 1,4-naphthoquinone (1.5 μM) was used as a positive

control for the inhibition of the respiratory chain, whereas 0.1% TX-100 provided the control

for full plasma membrane permeabilization139 (Figure 19, panels A and B).

Page 135: Protein-kinase inhibitors as potential leishmanicidal drugs

120

Figure 18. Luminescence principle of 3-LUC L. donovani promastigotes. In the presence of

DMNPE-luciferin, the cytoplasmic luciferase expressed by 3-LUC L. donovani promastigotes produces

luminescence with consumption of intracellular ATP. Depletion of the intracellular ATP levels, caused

by either plasma membrane permeabilization or inhibition of mitochondrial of ATP synthesis, cuts out

the luminescence of this strain.

In order to rule out plasma membrane permeabilization as the cause of a decrease in

ATP levels, entrance of SYTOX Green into L. infantum promastigotes treated with the

chosen compounds was monitored. SYTOX Green is a cationic vital dye with poor

fluorescence in aqueous media that greatly increases after its intercalation in DNA.141 Since

SYTOX Green is non-permeable to an unscathed membrane (MW = 600), an increase in its

fluorescence can only occur after the severe disruption of the plasma membrane of the

parasite. As in the 3-LUC assay, 0.1% TX-100 was used as a positive control for full

permeabilization of the parasite (Figure 19, panels C and D).

For both assays, TDZDs 1 and 2, HMKs 6, 28, 29 and 32, and quinones 72, 75-80 were

assayed at their respective IC80; an equipotent concentration aimed to produce a large but

selective effect, as well as to facilitate comparison among the compounds. Compound 33

was excluded due to its nil leishmanicidal activity on the Leishmania promastigote.

Page 136: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

121

Time (min)

0 10 20 30

0

20

40

60

80

100

0 10 20 30

72 (3 M)

75 (2 M)

76 (2 M)

77 (6 M)

78 (4.3 M)

79 (8 M)

80 (7.5 M)

Untreated0.1% TX-100 0.1% TX-100

0

20

40

60

80

100

0.1% TX-100

1.5 M1,4-naphtoquinone

1 (20 M)

2 (22 M)

6 (10 M)

28 (20 M)

29 (6.2 M)

32 (10 M)

Lum

ines

cence

(% r

espec

t to

the

contr

ol)

SY

TO

X G

reen

flu

ore

scen

ce(%

res

pec

t to

0.1

% T

X-1

00)

A B

C D

Figure 19. Assessment of intracellular ATP variation and plasma membrane permeabilization by

the compounds selected. Panels A and B: Real-time monitoring of the luminescence of 3-LUC

L. donovani promastigotes after addition (t = 0) of the selected TDZDs, HMKs (Panel A) and quinones

(Panel B) at their respective IC80. Luminescence values were represented as percentage with respect to

the control of untreated parasites. 1,4-naphthoquinone (1.5 µM) and Triton X-100 (0.1%) were added as

controls for inhibition of oxidative phosphorylation and plasma membrane permeabilization respectively.

Panels C and D: Variation of SYTOX Green fluorescence after addition (t = 0) of the selected TDZDs,

HMKs (Panel C) and quinones (Panel D) to L. infantum promastigotes. Fluorescence settings:

λEX = 485 nm, λEM = 520 nm. Arrow: addition of 0.1% Triton X-100 (final concentration).

The quinones 72 (3 µM), 75 (2 µM) and 76 (2 µM) caused a rapid and irreversible

inhibition of the luminescence of 3-LUC L. donovani promastigotes, at a similar or even

more pronounced rate than the control treatment with 1,4-naphthoquinone at 1.5 µM (Figure

19, panel B). The decrease in luminescence caused by the HMK 6 (10 µM) was also quite

similar to 1.5 µM 1,4-naphthoquinone, while its analogue 32 (10 µM) barely induced a 20%

decrease in luminescence. Notably, the TDZD 1 induced a sharp but transitory drop in

luminescence that reverted to initial values by 10 minutes (Figure 19, panel A). None of

these compounds showed an increase of SYTOX Green fluorescence within the monitoring

Page 137: Protein-kinase inhibitors as potential leishmanicidal drugs

122

time range (Figure 19, panels C and D). Consequently, membrane permeabilization was

ruled out as the mechanism of action of the compounds that caused the bioenergetic collapse

of the 3-LUC L. donovani promastigotes.

Once the severe disruption of the plasma membrane integrity was dismissed, the other

feasible target of the aforementioned compounds was the mitochondrial ATP synthesis. As

for compound 1, this TDZD most likely induces a transient leak of cytoplasmic ATP that we

may surmise to be due to a brief plasma membrane depolarization of the parasite.

4.4.1.2. Inhibition of the electrochemical potential of the Leishmania mitochondrion (ΔΨm)

In Leishmania, especially on the promastigote, oxidative phosphorylation is the main

source for ATP biosynthesis.66 This process is coupled to the respiratory chain with

generation of an electrochemical potential across the inner membrane of the mitochondrion

(ΔΨm). In fact, mitochondrial membrane depolarization is associated with inhibition of

oxidative phosphorylation, as well as activation of the intrinsic apoptotic pathway.141

The ΔΨm is the highest among the intracellular organelles, with a reported value

of -160 mV in L. major promastigotes.212 This accounts for the preferential accumulation of

hydrophobic cations into the mitochondrion under the Nernst equation.213 Hence, the

variation of the ΔΨm of Leishmania parasites can be easily monitored through the differential

accumulation of rhodamine 123, a fluorescent hydrophobic cation.141 The HMK 6 and the

quinones 72, 75 and 76 were selected from the previous assays to test their effect on the ΔΨm

of L. infantum promastigotes. The TDZD 1 was also selected in order to completely rule out

the mitochondrion of the parasite as its off-target. Compounds were incubated for 4 h with

the parasites at their respective IC80s prior to rhodamine 123 accumulation (described in

Materials and Methods). A 40 min incubation with 20 mM KCN was used as a control

inhibitor of oxidative phosphorylation (Figure 20).

Page 138: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

123

Figure 20. Rhodamine 123 accumulation in L. infantum promastigotes and its variation by the

selected compounds at their IC80. The number codes of the compounds are at the right top corner of

their respective panel. Parasites were incubated with the compounds as described in the Materials and

Methods section, followed by incubation with rhodamine 123. Rhodamine 123 accumulation was

measured by flow cytometry (λEX = 488 nm, λEM = 520 nm). Fluorescence values were expressed as a

percentage value with respect to the untreated control. Incubation with 20 mM KCN for 40 min was used

as positive control for mitochondrial membrane depolarization.

According to Figure 20, all the tested compounds decreased the accumulation of

rhodamine 123 in L. donovani promastigotes except for TDZD 1, as expected from its lack

of effect on the intracellular ATP content of 3-LUC L. donovani promastigotes. Compounds

6 (10 μM) and 72 (3 μM) induced the highest decrease of rhodamine 123 accumulation,

similar to the KCN-treated control (ca 70 %). Quinones 75 and 76 at 2 μM showed similar

effects, with 33 and 26.4 % decrease of rhodamine 123 fluorescence respectively.

Page 139: Protein-kinase inhibitors as potential leishmanicidal drugs

124

Since lower rhodamine 123 accumulation is caused by mitochondrial membrane

depolarization, these results confirmed that HMK 6 and quinones 72, 75 and 76 inhibit the

oxidative phosphorylation of L. donovani.

For quinones 72¸ 75 and 76, their resemblance to ubiquinone is a feasible hypothesis

for their inhibitory activity on the respiratory chain of the parasite. Ubiquinone acts as a

membrane bound electron carrier among complexes I, II and III. The molecule is associated

with the inner mitochondrial membrane through its polyisoprenoid chain while its

benzoquinone ring accepts the electrons from complex I and II, and transfers them to

complex III.214 As the activity of complex I in trypanosomatids is rather poor,71 complex II

and III remain the most feasible targets for these compounds, as reported for other

quinones.191,215,216

4.4.1.3. Inhibition of O2 consumption

As mentioned before, electron transport throughout the respiratory chain is coupled to

mitochondrial ATP synthesis. Hence, inhibition of one of these processes results in the

inhibition of the other and vice versa. Consequently, the mitochondrial depolarizing

compounds 6 (Figure 22), 72, 75 and 76 (Figure 23) were assayed for inhibition of the

respiration of Leishmania promastigotes by polarographic monitoring using a Clark oxygen

electrode.

The Clark oxygen electrode consists of a silver anode and a platinum cathode

connected with a sealed and thermostatised jacketed-chamber through a Teflon membrane.

A magnetic stirrer ensures the rapid equilibration of the O2 content throughout the whole

fluid (Figure 21). The Teflon membrane allows diffusion of oxygen from the respiration

buffer within the chamber to the electrodes, where an electrochemical reaction creates a

voltage proportional to the oxygen concentration.217 The cellular density of the parasite

suspension for Clark electrode experiments was 5-fold higher with respect to the standard

one used for the other short-term functional experiments. This approach allowed speeding

up the time threshold for the observation of effects on parasite respiration before the O2

available was fully consumed.

Page 140: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

125

Figure 21. Clark-type oxygen electrode scheme.218 This device allows for the monitoring of O2

consumption of a cellular solution placed in the incubation chamber. The chamber is completely sealed

off during the experiment to avoid reconstitution of O2 levels. The magnetic stirrer bar at the bottom of

the chamber ensures homogeneous O2 levels across the cellular solution. Water at the needed temperature

flows through the “jacket” surrounding the incubation chamber to maintain the appropriate temperature

during the experiment. The electrodes that record the electrochemical changes in the solution associated

with the O2 levels are placed in a separate chamber in KCl solution, connected to the incubation chamber

through a Teflon membrane.

The halomethylketone 6 inhibited the respiration of L. infantum promastigotes in a

concentration dependent manner (Figure 22, panels B and C), reaching up to 51% inhibition

at 2-fold (20 μM) and 63% inhibition at 5-fold (50 μM) its IC80.

Page 141: Protein-kinase inhibitors as potential leishmanicidal drugs

126

Figure 22. Oxygen consumption rates of living L. infantum parasites treated with the HMK 6.

Parasites were prepared in Respiration Buffer at 5-fold the usual cellular concentration established for

short-termed experiments (108 parasites/mL), and maintained at 26 ºC. Panel A: Untreated parasites.

Panels B and C: Parasites treated with HMK 6 at 2- and 5-fold its IC80 respectively. Values represent the

percentage (%) of O2 consumption rate. The different colours of the dotted lines account for the slope

measurements in each step of the experiment. The use of high concentrations of parasites and compounds

allowed for the observation of the effects of the compounds before O2 levels were fully depleted.

On the other hand, the L. infantum promastigotes treated with quinones 72, 75 and 76

at 3.3- to 5-fold their IC80 (10 μM) showed a double effect: the quinones initially induced an

acceleration of the O2 consumption rates, but the respiration of the parasites slowed down

afterwards (Figure 23, panels B, C and D). A feasible explanation is that the initial

acceleration entails a compensation mechanism to recover from the decrease in intracellular

ATP levels caused by the inhibition of oxidative phosphorylation by the quinones. Over

time, this increase is no longer feasible and the parasites undergo a steady loss of respiration

that also may be related to direct inhibition of the respiratory chain by the compounds.

Page 142: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

127

Figure 23. Oxygen consumption rates of L. infantum parasites with quinones 72, 75 and 76. The

addition of each compound is represented with a red arrow and its code number in bold black. Panel A:

untreated parasites. Panel B: treatment with compounds 72 at 3.3-fold its IC80. Panel C: treatment with

compound 75 at 5-fold its IC80. Panel D: treatment with compound 76 at 5-fold its IC80. Values represent

the percentage of O2 consumption rate with respect to the basal rate, before the addition of the compounds.

The different colours of the dotted lines account for the different slopes, whose respective percentage

values appear at the lower left corner of each panel.

4.4.1.4. Identification of the PKI target inside the respiratory chain

In Leishmania, since complex I is barely active,71 complex II is the main electron donor

for complex III through ubiquinone using succinate as substrate. Consequently, the proton

gradient across the membrane depends on complexes III and IV, with O2 as the final electron

acceptor at complex IV. The F0F1-ATPase (complex V) carries out ATP synthesis by

oxidative phosphorylation profiting from the H+ gradient created across the inner

membrane.63 Under this scenario, the site of inhibition for each compound was mapped by

using specific substrates and inhibitors for each complex on the mitochondrion of digitonin-

permeabilized L. infantum promastigotes.

Page 143: Protein-kinase inhibitors as potential leishmanicidal drugs

128

The isolation of the single mitochondrion of trypanosomatids in an intact state is

unfeasible due to the large size of this organelle (ca 12% of the intracellular volume of the

parasite), along with the extreme resistance to mechanical disruption of the plasma

membrane due to its subpellicular layer of microtubules. This issue is solved by the use of

digitonin, a non-ionic detergent. At low concentrations, digitonin selectively disrupts the

plasma membrane through a preferential interaction with sterols, which are absent in the

inner mitochondrial membrane. This selective permeation of the plasma membrane affords

direct access of respiratory substrates and inhibitors to a functional mitochondrion with an

intact inner membrane.219

Figure 24 provides an explanation for this strategy. Panel A shows the variation of

control parasites after digitonin addition. The respiration increases after ADP addition,

accounting for the coupling degree between the respiratory chain and ATP synthesis.

Addition of FCCP, a protonophore molecule that dissipates the H+ gradient across the inner

membrane, ensures the maximal speed of the respiratory chain. Panel A’ provides a graphical

explanation of the polarographic pattern. Panel B shows the inhibition of respiration through

the inhibition of F0F1-ATPase by oligomycin (panel B’1), and the uncoupling of this

complex from the respiratory chain by FCCP addition (panel B’2).

Page 144: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

129

Figure 24. Oxygen consumption rates of digitonin-permeabilized L. infantum parasites. Panel A:

FCCP-treated mitochondria as a control of uncoupled respiration. Panel B: Control of uncoupled

respiration with oligomycin-inhibited F0F1-ATP synthase. Panels A’, B’1 and B’2: schemes of the effects

of FCCP and oligomycin on the respiratory chain of Leishmania. Final concentrations of substrates and

inhibitors: 60 μM digitonin (Dig), 100 μM ADP, 10 μM FCCP, 12.6 μM oligomycin (Omy).

Addition of TMPD-ascorbate, an electron donor for cytochrome c oxidase (complex

IV) led to a partial recovery of the respiration inhibited by HMK 6. This excluded complexes

IV and V as targets for this compound (Figure 25, panels A and A’). When the respiratory

chain of Leishmania is inhibited at complex II by malonate, the single source to feed

electrons into complex III is through complex I by addition of its substrate,

Page 145: Protein-kinase inhibitors as potential leishmanicidal drugs

130

α-glycerophosphate (Figure 25, panels B and B’). As α-glycerophosphate did not improve

the respiration rate inhibited by compound 6, complex II was discarded as a target, leaving

complex III as the main inhibition site within the respiratory chain for this halomethylketone

(Figure 25, panels C and C’). Remarkably, the alkaline pH of the mitochondrion favours the

formation of thiolate, which increases the reactivity of nucleophilic compounds such as

halomethylketones.

The importance of cytochrome c reductase (complex III) as a leishmanicidal target is

endorsed by leishmanicidal compounds acting on this complex, including antimycin A and

other compounds in clinical use for other diseases caused by apicomplexans, such as

buparvaquone (hydroxynaphthoquinone) for theileriosis and tafenoquine

(8-aminoquinoline) for malaria.191,216 Other compounds of natural origin that act on the

cytochrome c reductase of Leishmania include chromanol derivatives, which are related to

vitamin E, and essential oil components, among others.220,221

Page 146: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

131

Figure 25. Oxygen consumption rates of digitonized-L. infantum promastigotes treated with HMK

6 and the complex II-inhibitor malonate. ADP addition prevents its potential role as a limiting substrate

for ATP synthesis by complex V. Panel A: Variation of the respiration rate by addition of TMPD-Asc on

promastigotes treated with 50 μM 6. Panel B: Variation of respiration after addition of α-GP under

inhibition of complex II. Panel C: Inability of α-glycerophosphate to restore the inhibition of the

respiration rate induced by compound 6. Panels A’, B’ and C’: Graphical interpretation of the effects of

HMK 6, malonate and α-glycerophosphate on respiration from the respective panels on their left (A, B

and C). Final concentration for substrates and inhibitors: 60 μM digitonin (Dig), 100 μM ADP, 10 mM

malonate, 6.7 mM α-glycerophosphate (α-GP), TMPD-ascorbate (TMPD-Asc) 0.1 mM and 1.7 mM,

respectively.

Page 147: Protein-kinase inhibitors as potential leishmanicidal drugs

132

As the addition of quinones 72, 75 and 76 increases respiration rates on intact parasites

(Figure 23), a different approach was taken. Under inhibition of complex V by oligomycin,

the addition of any of these three quinones increased the respiration rate (Figure 26,

panels A, B and C), as seen for the protonophore FCCP in Figure 24. These results confirmed

a mild uncoupling effect of these compounds. Nevertheless, a dual action of these quinones

was anticipated. In fact, these three quinones reduced the oxygen consumption rate of

FCCP-treated mitochondria, demonstrating inhibitory activity on the electron transport chain

of the parasite (Figure 26 D, E and F). Inhibition of respiration by the quinones was restored

after addition of TMPD-Asc, therefore the target of inhibition of compounds 72, 75 and 76

is upstream complex IV (Figure 26 G, H and I). Nevertheless, the acceleration of respiration

in living parasites induced by these quinones (Figure 23) points towards a higher uncoupling

than inhibitory effect in vivo.

Page 148: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

133

Figure 26. Oxygen consumption rates of digitonin-permeabilized L. infantum parasites. Panels A, B

and C: Uncoupling effect of compounds 72, 75 and 76 (respectively) in oligomycin-treated

mitochondria. Panels D, E and F: Inhibition of respiration by compounds 72, 75 and 76 (respectively) in

FCCP-uncoupled mitochondria. Panels G, H and I: Restoration of the respiration inhibited by compounds

72, 75 and 76 (respectively) after addition of TMPD-ascorbate. Final concentrations of substrates and

inhibitors: 60 μM digitonin (Dig), 100 μM ADP, 10 μM FCCP, 12.6 μM oligomycin (Omy),

TMPD-ascorbate (TMPD-Asc) 0.1 mM and 1.7 mM respectively.

Page 149: Protein-kinase inhibitors as potential leishmanicidal drugs

134

4.4.1.5. Induction of programmed cell death in L. infantum promastigotes

Most of the morphological features of apoptosis in mammalian cells, such a chromatin

condensation, shrinkage of the cytoplasm, or blebbing of the plasma membrane are

reproduced in Leishmania, but the molecular executioners are poorly defined in the

parasite.222 In trypanosomatids, the only feasible apoptotic pathway is the intrinsic one, as

the external apoptotic pathway is absent.223 The role of the mitochondrion in the intrinsic

pathway of Leishmania is mandatory; its depolarization is an early apoptotic stage, followed

by chromatin degradation at later stages with increase of the subG0/G1 population in cell

cycle studies.141 Hence, we evaluated the induction of programmed cell death in Leishmania

promastigotes with the mitochondrial depolarizing compounds 6, 72, 75 and 76. Other

leishmanicidal drugs with known apoptotic-inducing effects linked to inhibition of the

respiratory chain of Leishmania include miltefosine, tafenoquine and sitamaquine, targeting

complex IV, III and II of the respiratory chain, respectively.215,216,224 Pentamidine and

paromomycin also induce programmed cell death by targeting the mitochondrion but their

target remains unveiled.127,131

The individual content of DNA in a population of L. infantum promastigotes was

analysed by flow cytometry with propidium iodide (PI). PI is a fluorescent probe that

stoichiometrically binds to DNA. Since DNA fragmentation reduces the capacity of DNA to

bind PI, apoptotic cells are found in the SubG0/G1 population, with a much lower PI

fluorescence with respect to other cell populations in different phases of the cell cycle

(G0/G1, S, G2/M).141 For this study, compounds 6, 72, 75 and 76 were assayed at their

respective IC80. Miltefosine at 15 μM was used as a control to obtain an apoptotic

histogram.225

Both the HMK and the quinones tested increased the SubG0/G1 population, with lower

levels of PI accumulation in apoptotic parasites due to chromatin degradation (Figure 27).

Hence, it was concluded that HMK 6 and quinones 72, 75 and 75 induced programmed cell

death in L. infantum promastigotes.

Page 150: Protein-kinase inhibitors as potential leishmanicidal drugs

Results and discussion

135

Figure 27. Cell cycle of L. infantum promastigotes treated with quinones 72, 75 and 76 and the

halomethylketone 6 at their IC80. The values represented are the percentage of SubG0/G1 population

(delimited inside green bars) as the typical population for apoptotic-like process (degraded chromatin).

Miltefosine (HePc) at 15 μM was added as a control for induction of apoptosis.

Altogether, one halomethylketone and three naphthoquinones out of the 14

leishmanicidal LdGSK-3s inhibitors from our in-house chemical library induced a

bioenergetic collapse in the promastigote of Leishmania by targeting the respiratory chain

of the parasite, leading to apoptosis and revealing the multi-target nature of these protein

kinase inhibitors.

Page 151: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 152: Protein-kinase inhibitors as potential leishmanicidal drugs

CONCLUSIONS

Page 153: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 154: Protein-kinase inhibitors as potential leishmanicidal drugs

Conclusions

139

5. Conclusions

In this PhD thesis, new leishmanicidal agents belonging to three different chemical

families acting on Leishmania GSK-3s at micromolar concentrations have been identified

and characterized. The energy metabolism of Leishmania has been identified as an additional

off-target mechanism for members of two different chemical scaffolds.

From the results obtained, the following conclusions were drawn:

1. The usefulness of collections of small-size organic compounds as a source for the

identification of new leishmanicidal inhibitors for LdGSK-3 has been endorsed.

2. The combination of a target-based approach with an unbiased phenotypic assay is

highly relevant for a full interrogation of a chemical library in search of effective

compounds with a known mechanism of action.

3. The GSK-3s of Leishmania was confirmed as a druggable target by the identification

of 16 active compounds from four different chemical families; thiadiazolidindiones,

iminothiadiazoles, halomethylketones and quinones, broadening the range of

chemical scaffolds for inhibition of this enzyme.

4. A selectivity index over 10, a major requirement to endorse further hit optimization,

was obtained for TDZD 1, HMK 28 and quinones 75, 76 and 79.

5. The halomethylketones 6 and 32, and the quinones 72, 76 and 79, decreased the

parasite load in infected murine peritoneal macrophages.

6. Among the 24 1,4-naphthoquinones tested, simultaneous inhibition of both

Leishmania proliferation and LdGSK3-s activity took place in scaffolds with a

Page 155: Protein-kinase inhibitors as potential leishmanicidal drugs

140

carbamate substituent at the nitrogen of the 2-amino group and a Cl atom at

position 3. Moreover, quinones with this pattern of substitution are the first selective

LdGSK-3s inhibitors with no activity against HsGSK-3β described so far. Hence,

this chemical pattern is the most appealing for further optimization.

7. The energy metabolism of the promastigote is a useful parameter to gauge its

intracellular homeostasis, and find feasible off-targets for GSK-3 inhibitors. Besides

inhibition of LdGSK-3s, the HMK 6 and the quinones 72, 75 and 76 caused a

bioenergetic collapse in Leishmania promastigotes, with mitochondrial

depolarization and inhibition of both respiration and ATP synthesis, driving the

parasite into a programmed cell death.

8. Inside the respiratory chain of Leishmania promastigotes, complex III was identified

as a target for HMK 6, whereas a double effect was observed for quinones 72, 75 and

76, with uncoupling and inhibitory activities on the respiration of the parasites.

In conclusion, the combination of a target-based screening focused on the GSK-3s of

Leishmania with a phenotypic screening in the parasite has led to the identification of new

compounds with appealing activities for their further development as new leishmanicidal

drugs. Moreover, some of these compounds have an additional target in the respiratory chain

of the parasite. The multi-target nature of these compounds is an advantageous trait for a

jeopardized raise of resistance caused by mutation in a single target.

Page 156: Protein-kinase inhibitors as potential leishmanicidal drugs

BIBLIOGRAPHY

Page 157: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 158: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

143

Bibliography

1. Duthie, M. S.; Lison, A.; Courtenay, O., Advances toward diagnostic tools for managing

zoonotic visceral leishmaniasis. Trends Parasitol, 2018, 34, 881-890.

2. Salotra, P.; Kaushal, H.; Ramesh, V., Containing post Kala-Azar Dermal Leishmaniasis

(PKDL): Pre-requisite for sustainable elimination of visceral leishmaniasis (VL) from South Asia in

Kala Azar in South Asia, ed. E. Noiri; T. K. Jha, Springer International Publishing, 2017, pp. 7-21.

3. World Health Organization. Global Health Observatory (GHO) data. Leishmaniasis.

https://www.who.int/gho/neglected_diseases/leishmaniasis/en/. Accessed September 2020.

4. Babayan, S. A.; Schneider, D. S., Immunity in society: Diverse solutions to common problems.

PloS Biol, 2012, 10, e1001297.

5. González, E.; Jiménez, M.; Hernández, S.; Martín-Martín, I.; Molina, R., Phlebotomine sand

fly survey in the focus of leishmaniasis in Madrid, Spain (2012-2014): Seasonal dynamics,

Leishmania infantum infection rates and blood meal preferences. Parasites Vectors, 2017, 10, 368.

6. Carrillo, E.; Moreno, J.; Cruz, I., What is responsible for a large and unusual outbreak of

leishmaniasis in Madrid? Trends Parasitol, 2013, 29, 579-580.

7. Cesinaro, A. M.; Nosseir, S.; Mataca, E.; Mengoli, M. C.; Cavatorta, C.; Gennari, W., An

outbreak of cutaneous leishmaniasis in Modena province (Northern Italy): Report of 35 cases.

Pathologica, 2017, 109, 363-367.

8. Burza, S.; Croft, S. L.; Boelaert, M., Leishmaniasis. Lancet, 2018, 392, 951-970.

9. Hay, S. I.; Abajobir, A. A.; Abate, K. H.; Abbafati, C.; Abbas, K. M.; Abd-Allah, F.;

Abdulkader, R. S.; et al., Global, regional, and national disability-adjusted life-years (DALYs) for

333 diseases and injuries and healthy life expectancy (HALE) for 195 countries and territories, 1990-

2016: a systematic analysis for the Global Burden of Disease Study 2016. Lancet, 2017, 390, 1260-

1344.

10. Baileyid, F.; Mondragon-Shem, K.; Haines, L. R.; Olabi, A.; Alorfi, A.; Ruiz-Postigo, J. A.;

Alvar, J.; Hotez, P.; Adams, E. R.; Vélez, I. D.; Al-Salem, W.; Eaton, J.; Acosta-Serrano, Á.;

Page 159: Protein-kinase inhibitors as potential leishmanicidal drugs

144

Molyneux, D. H., Cutaneous leishmaniasis and co-morbid major depressive disorder: A systematic

review with burden estimates. PLoS Negl Trop Dis, 2019, 13, e0007092.

11. Feleke, B. E., Nutritional status of visceral leishmaniasis patients: A comparative cross-

sectional study. Clin Nutr ESPEN, 2019, 33, 139-142.

12. Martínez, D. Y.; Verdonck, K.; Kaye, P. M.; Adaui, V.; Polman, K.; Llanos-Cuentas, A.;

Dujardin, J. C.; Boelaert, M., Tegumentary leishmaniasis and coinfections other than HIV. PLoS

Negl Trop Dis, 2018, 12, e0006125.

13. Lindoso, J. A. L.; Cunha, M. A.; Queiroz, I. T.; Moreira, C. H. V., Leishmaniasis–HIV

coinfection: Current challenges. HIV/AIDS - Res Palliative Care, 2016, 8, 147-156.

14. Guarneri, C.; Tchernev, G.; Bevelacqua, V.; Lotti, T.; Nunnari, G., The unwelcome trio: HIV

plus cutaneous and visceral leishmaniasis. Dermatol Ther, 2016, 29, 88-91.

15. Colasanti, J.; Altamirano, J.; Espinoza, L., An unwelcome synergy: Leishmaniasis and HIV.

Am J Med, 2013, 126, 114-116.

16. Solomon, M.; Sahar, N.; Pavlotzky, F.; Barzilai, A.; Jaffe, C. L.; Nasereddin, A.; Schwartz,

E., Mucosal leishmaniasis in travelers with leishmania braziliensis complex returning to israel.

Emerg Infect Dis, 2019, 25, 642-648.

17. Schäfer, I.; Volkmann, M.; Beelitz, P.; Merle, R.; Müller, E.; Kohn, B., Retrospective

evaluation of vector-borne infections in dogs imported from the Mediterranean region and

southeastern Europe (2007-2015). Parasites Vectors, 2019, 12, 30.

18. Söbirk, S. K.; Inghammar, M.; Collin, M.; Davidsson, L., Imported leishmaniasis in Sweden

1993-2016. Epidemiol Infect, 2018, 146, 1267-1274.

19. Alawieh, A.; Musharrafieh, U.; Jaber, A.; Berry, A.; Ghosn, N.; Bizri, A. R., Revisiting

leishmaniasis in the time of war: the Syrian conflict and the Lebanese outbreak. Int J Infect Dis,

2014, 29, 115-119.

20. Jessen, L.; Bard, B.; Grenon, X.; Bally, F., Leishmaniasis: An infection of travellers and

migrants. Rev Med Suisse, 2016, 12, 1718-1722.

21. Booth, M., Climate change and the Neglected Tropical Diseases. Adv Parasitol, 2018; 100,

39-126.

Page 160: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

145

22. Short, E. E.; Caminade, C.; Thomas, B. N., Climate change contribution to the emergence or

re-emergence of parasitic diseases. Infec Dis, 2017, 10, 1178633617732296.

23. Jha, R. K.; Sah, A. K.; Shah, D. K.; Sah, P., The treatment of visceral leishmaniasis: Safety

and efficacy. J Nepal Med Assoc, 2013, 52, 645-651.

24. Okwor, I.; Uzonna, J., Social and economic burden of human leishmaniasis. Am J Trop Med

Hyg, 2016, 94, 489-493.

25. Ponte-Sucre, A.; Gamarro, F.; Dujardin, J. C.; Barrett, M. P.; López-Vélez, R.; García-

Hernández, R.; Pountain, A. W.; Mwenechanya, R.; Papadopoulou, B., Drug resistance and treatment

failure in leishmaniasis: A 21st century challenge. PLoS Negl Trop Dis, 2017, 11, e0006052.

26. Moafi, M.; Sherkat, R.; Taleban, R.; Rezvan, H., Leishmania vaccines entered in clinical trials:

A review of literature. Int J Prev Med, 2019, 10, 1-6.

27. Bhattacharya, P.; Ali, N., Treatment of visceral leishmaniasis: anomalous pricing and

distribution of AmBisome and emergence of an indigenous liposomal amphotericin B,

FUNGISOME. J Parasitic Dis, 2016, 40, 1094-1095.

28. Chatelain, E.; Ioset, J. R., Drug discovery and development for neglected diseases: the DNDi

model. Drug Des Devel Ther, 2011, 5, 175-181.

29. Schönian, G.; Lukeš, J.; Stark, O.; Cotton, J. A., Molecular evolution and phylogeny of

Leishmania in Drug resistance in Leishmania parasites, ed. A. Ponte-Sucre; M. Padrón-Nieves,

Springer, Cham, 2018, pp. 19-57.

30. Gramiccia, M.; Di Muccio, T., Diagnosis in The Leishmaniases: Old Neglected Tropical

Diseases, ed. F. Bruschi; L. Gradoni, Springer International Publishing, New York (NY), 2018, pp.

137-168.

31. van der Auwera, G.; Dujardina, J. C., Species typing in dermal leishmaniasis. Clin Microbiol

Rev, 2015, 28, 265-294.

32. Paul, S.; Pervin, M., Pathology and mechanism of disease in Kala-Azar, in Kala Azar in South

Asia, ed. E. Noiri; T. K. Jha, Springer International Publishing, 2017, pp. 3-6.

33. Tiwari, N.; Kishore, D.; Bajpai, S.; Singh, R., Visceral leishmaniasis: An immunological

viewpoint on asymptomatic infections and post kala azar dermal leishmaniasis. Asian Pac J Trop

Med, 2018, 11, 98-108.

Page 161: Protein-kinase inhibitors as potential leishmanicidal drugs

146

34. Zijlstra, E. E.; Musa, A. M.; Khalil, E. A. G.; El Hassan, I. M.; El-Hassan, A. M., Post-kala-

azar dermal leishmaniasis. Lancet Infect Dis, 2003, 3, 87-98.

35. Scorza, B. M.; Carvalho, E. M.; Wilson, M. E., Cutaneous manifestations of human and

murine leishmaniasis. Int J Mol Sci, 2017, 18, 1296.

36. Handler, M. Z.; Patel, P. A.; Kapila, R.; Al-Qubati, Y.; Schwartz, R. A., Cutaneous and

mucocutaneous leishmaniasis: Differential diagnosis, diagnosis, histopathology, and management. J

Am Acad Dermatol, 2015, 73, 911-926.

37. World Health Organization, Control of the leishmaniases: Report of a meeting of the WHO

Expert Committee on the control of leishmaniases, Geneva, 22-26 March 2010, ed., World Health

Organization, Geneva (Switzerland), 2010.

38. Zerpa, O.; Padrón-Nieves, M.; Ponte-Sucre, A., American tegumentary leishmaniasis in Drug

resistance in Leishmania parasites, ed. A. Ponte-Sucre; M. Padrón-Nieves, Springer International

Publishing, 2018, pp. 177-191.

39. Kaufer, A.; Ellis, J.; Stark, D.; Barratt, J., The evolution of trypanosomatid taxonomy. Parasite

Vectors, 2017, 10, 287.

40. Rosenzweig, D.; Smith, D.; Opperdoes, F.; Stern, S.; Olafson, R. W.; Zilberstein, D.,

Retooling Leishmania metabolism: from sand fly gut to human macrophage. FASEB J, 2008, 22,

590-602.

41. Chakraborty, B.; Biswas, S.; Mondal, S.; Bera, T., Stage specific developmental changes in

the mitochondrial and surface membrane associated redox systems of Leishmania donovani

promastigote and amastigote. Biochemistry (Mosc), 2010, 75, 494-518.

42. Hurrell, B. P.; Regli, I. B.; Tacchini-Cottier, F., Different Leishmania species drive distinct

neutrophil functions. Trends Parasitol, 2016, 32, 392-401.

43. Ready, P. D., Biology of phlebotomine sand flies as vectors of disease agents. Annu Rev

Entomol, 2013, 58, 227-250.

44. Rogers, M. E.; Chance, M. L.; Bates, P. A., The role of promastigote secretory gel in the origin

and transmission of the infective stage of Leishmania mexicana by the sandfly Lutzomyia

longipalpis. Parasitology, 2002, 124, 495-507.

Page 162: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

147

45. de Menezes, J. P.; Saraiva, E. M.; da Rocha-Azevedo, B., The site of the bite: Leishmania

interaction with macrophages, neutrophils and the extracellular matrix in the dermis. Parasite

Vectors, 2016, 9, 264.

46. Antoine, J. C.; Prina, E.; Courret, N.; Lang, T., Leishmania spp.: on the interactions they

establish with antigen-presenting cells of their mammalian hosts. Adv Parasitol, 2004, 58, 1-68.

47. De Pablos, L. M.; Ferreira, T. R.; Walrad, P. B., Developmental differentiation in Leishmania

lifecycle progression: post-transcriptional control conducts the orchestra. Curr Opin Microbiol,

2016, 34, 82-89.

48. Kamhawi, S., The yin and yang of leishmaniasis control. PLoS Negl Trop Dis, 2017, 11,

e0005529.

49. Uliana, S. R. B.; Trinconi, C. T.; Coelho, A. C., Chemotherapy of leishmaniasis: present

challenges. Parasitology, 2018, 145, 464-480.

50. Coutinho De Oliveira, B.; Duthie, M. S.; Alves Pereira, V. R., Vaccines for leishmaniasis and

the implications of their development for American tegumentary leishmaniasis. Hum Vaccin

Immunother, 2019, 1-12.

51. Moafi, M.; Rezvan, H.; Sherkat, R.; Taleban, R., Leishmania vaccines entered in clinical trials:

A review of literature. Int J Prev Med, 2019, 10, 95.

52. Selvapandiyan, A.; Croft, S. L.; Rijal, S.; Nakhasi, H. L.; Ganguly, N. K., Innovations for the

elimination and control of visceral leishmaniasis. PLoS Negl Trop Dis, 2019, 13, e0007616.

53. Claborn, D. M., The biology and control of leishmaniasis vectors. J Glob Infect Dis, 2010, 2,

127-134.

54. Gonzalez, U.; Pinart, M.; Sinclair, D.; Firooz, A.; Enk, C.; Velez, I. D.; Esterhuizen, T. M.;

Tristan, M.; Alvar, J., Vector and reservoir control for preventing leishmaniasis. Cochrane Database

Syst Rev, 2015, CD008736.

55. Seva, A. D. P.; Martcheva, M.; Tuncer, N.; Fontana, I.; Carrillo, E.; Moreno, J.; Keesling, J.,

Efficacies of prevention and control measures applied during an outbreak in Southwest Madrid,

Spain. PLoS One, 2017, 12, e0186372.

Page 163: Protein-kinase inhibitors as potential leishmanicidal drugs

148

56. Amman, J., Report on the interregional meeting on leishmaniasis among neighbouring

endemic countries in the Eastern Mediterranean, African and European regions. WHO Regional

Office for the Eastern Mediterranean, Cairo (Egypt), 2019.

57. Gradoni, L.; López-Vélez, R.; Mokni, M., Manual on case management and surveillance of

the leishmaniases in the WHO European Region. WHO Regional Office for Europe, Copenhagen

(Denmark), 2017.

58. Ibarra-Meneses, A. V.; Moreno, J.; Carrillo, E., New strategies and biomarkers for the control

of visceral leishmaniasis. Trends Parasitol, 2020, 36, 29-38.

59. Dixit, K. K.; Verma, S.; Singh, O. P.; Singh, D.; Singh, A. P.; Gupta, R.; Negi, N. S.; Das, P.;

Sundar, S.; Singh, R.; Salotra, P., Validation of SYBR green I based closed tube loop mediated

isothermal amplification (LAMP) assay and simplified direct-blood-lysis (DBL)-LAMP assay for

diagnosis of visceral leishmaniasis (VL). PLoS Negl Trop Dis, 2018, 12, e0006922.

60. Menna-Barreto, R. F.; de Castro, S. L., The double-edged sword in pathogenic

trypanosomatids: The pivotal role of mitochondria in oxidative stress and bioenergetics. Biomed Res

Int, 2014, 2014, 614014.

61. Haanstra, J. R.; Gonzalez-Marcano, E. B.; Gualdron-Lopez, M.; Michels, P. A., Biogenesis,

maintenance and dynamics of glycosomes in trypanosomatid parasites. Biochim Biophys Acta, 2016,

1863, 1038-1048.

62. Docampo, R.; Moreno, S. N., Acidocalcisome: A novel Ca2+ storage compartment in

trypanosomatids and apicomplexan parasites. Parasitol Today, 1999, 15, 443-448.

63. Fidalgo, L. M.; Gille, L., Mitochondria and trypanosomatids: targets and drugs. Pharm Res,

2011, 28, 2758-2770.

64. de Souza, W.; Attias, M.; Rodrigues, J. C., Particularities of mitochondrial structure in

parasitic protists (Apicomplexa and Kinetoplastida). Int J Biochem Cell Biol, 2009, 41, 2069-2080.

65. McConville, M. J.; Saunders, E. C.; Kloehn, J.; Dagley, M. J., Leishmania carbon metabolism

in the macrophage phagolysosome- feast or famine? F1000Res, 2015, 4, 938.

66. Tielens, A. G.; van Hellemond, J. J., Surprising variety in energy metabolism within

Trypanosomatidae. Trends Parasitol, 2009, 25, 482-490.

Page 164: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

149

67. Saunders, E. C.; Ng, W. W.; Chambers, J. M.; Ng, M.; Naderer, T.; Kromer, J. O.; Likic, V.

A.; McConville, M. J., Isotopomer profiling of Leishmania mexicana promastigotes reveals

important roles for succinate fermentation and aspartate uptake in tricarboxylic acid cycle (TCA)

anaplerosis, glutamate synthesis, and growth. J Biol Chem, 2011, 286, 27706-27717.

68. Saunders, E. C.; Ng, W. W.; Kloehn, J.; Chambers, J. M.; Ng, M.; McConville, M. J., Induction

of a stringent metabolic response in intracellular stages of Leishmania mexicana leads to increased

dependence on mitochondrial metabolism. PLoS Pathog, 2014, 10, e1003888.

69. Sarnoff, R.; Desai, J.; Desjeux, P.; Mittal, A.; Topno, R.; Siddiqui, N. A.; Pandey, A.; Sur, D.;

Das, P., The economic impact of visceral leishmaniasis on rural households in one endemic district

of Bihar, India. Trop Med Int Health, 2010, 15, 42-49.

70. Teodoro, J. S.; Rolo, A. P.; Palmeira, C. M., The NAD ratio redox paradox: why does too

much reductive power cause oxidative stress? Toxicol Mech Methods, 2013, 23, 297-302.

71. Opperdoes, F. R.; Michels, P. A., Complex I of Trypanosomatidae: does it exist? Trends

Parasitol, 2008, 24, 310-317.

72. Duarte, M.; Tomas, A. M., The mitochondrial complex I of trypanosomatids--an overview of

current knowledge. J Bioenerg Biomembr, 2014, 46, 299-311.

73. Mondal, S.; Roy, J. J.; Bera, T., Generation of adenosine tri-phosphate in Leishmania donovani

amastigote forms. Acta Parasitol, 2014, 59, 11-16.

74. Naderer, T.; Heng, J.; McConville, M. J., Evidence that intracellular stages of Leishmania

major utilize amino sugars as a major carbon source. PLoS Pathog, 2010, 6, e1001245.

75. Eisele, N. A.; Ruby, T.; Jacobson, A.; Manzanillo, P. S.; Cox, J. S.; Lam, L.; Mukundan, L.;

Chawla, A.; Monack, D. M., Salmonella require the fatty acid regulator PPARdelta for the

establishment of a metabolic environment essential for long-term persistence. Cell Host Microbe,

2013, 14, 171-182.

76. Bera, T.; Lakshman, K.; Ghanteswari, D.; Pal, S.; Sudhahar, D.; Islam, M. N.; Bhuyan, N. R.;

Das, P., Characterization of the redox components of transplasma membrane electron transport

system from Leishmania donovani promastigotes. Biochim Biophys Acta, 2005, 1725, 314-326.

77. Biswas, S.; Haque, R.; Bhuyan, N. R.; Bera, T., Participation of chlorobiumquinone in the

transplasma membrane electron transport system of Leishmania donovani promastigote: effect of

Page 165: Protein-kinase inhibitors as potential leishmanicidal drugs

150

near-ultraviolet light on the redox reaction of plasma membrane. Biochim Biophys Acta, 2008, 1780,

116-127.

78. Vickers, T. J.; Beverley, S. M., Folate metabolic pathways in Leishmania. Essays Biochem,

2011, 51, 63-80.

79. McCall, L. I.; El Aroussi, A.; Choi, J. Y.; Vieira, D. F.; De Muylder, G.; Johnston, J. B.; Chen,

S.; Kellar, D.; Siqueira-Neto, J. L.; Roush, W. R.; Podust, L. M.; McKerrow, J. H., Targeting

ergosterol biosynthesis in Leishmania donovani: essentiality of sterol 14 alpha-demethylase. PLoS

Negl Trop Dis, 2015, 9, e0003588.

80. Singh, K.; Garg, G.; Ali, V., Current therapeutics, their problems and thiol metabolism as

potential drug targets in leishmaniasis. Curr Drug Metab, 2016, 17, 897-919.

81. Kumar, A.; Chauhan, N.; Singh, S., Understanding the cross-talk of redox metabolism and Fe-

S cluster biogenesis in Leishmania through systems biology approach. Front Cell Infect Microbiol,

2019, 9, 15.

82. Tagoe, D. N.; Kalejaiye, T. D.; de Koning, H. P., The ever unfolding story of cAMP signaling

in trypanosomatids: vive la difference! Front Pharmacol, 2015, 6, 185.

83. Manning, G.; Whyte, D. B.; Martinez, R.; Hunter, T.; Sudarsanam, S., The protein kinase

complement of the human genome. Science, 2002, 298, 1912-1934.

84. Fabbro, D.; Cowan-Jacob, S. W.; Moebitz, H., Ten things you should know about protein

kinases: IUPHAR Review 14. Br J Pharmacol, 2015, 172, 2675-2700.

85. Parsons, M.; Worthey, E. A.; Ward, P. N.; Mottram, J. C., Comparative analysis of the

kinomes of three pathogenic trypanosomatids: Leishmania major, Trypanosoma brucei and

Trypanosoma cruzi. BMC Genomics, 2005, 6, 127.

86. Borba, J. V. B.; Silva, A. C.; Ramos, P. I. P.; Grazzia, N.; Miguel, D. C.; Muratov, E. N.;

Furnham, N.; Andrade, C. H., Unveiling the kinomes of Leishmania infantum and L. braziliensis

empowers the discovery of new kinase targets and antileishmanial compounds. Comput Struct

Biotechnol J, 2019, 17, 352-361.

87. Cell Signaling Technology. Protein kinases: Interactive human kinome

https://www.cellsignal.com/contents/science-protein-kinases/protein-kinases-interactive-human-

kinome/kinases-human-kinome. Accessed September 2020.

Page 166: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

151

88. Morales, M. A.; Watanabe, R.; Laurent, C.; Lenormand, P.; Rousselle, J. C.; Namane, A.;

Spath, G. F., Phosphoproteomic analysis of Leishmania donovani pro- and amastigote stages.

Proteomics, 2008, 8, 350-363.

89. Tsigankov, P.; Gherardini, P. F.; Helmer-Citterich, M.; Spath, G. F.; Myler, P. J.; Zilberstein,

D., Regulation dynamics of Leishmania differentiation: Deconvoluting signals and identifying

phosphorylation trends. Mol Cell Proteomics, 2014, 13, 1787-1799.

90. Jones, N. G.; Catta-Preta, C. M. C.; Lima, A.; Mottram, J. C., Genetically validated drug

targets in Leishmania: Current knowledge and future prospects. ACS Infect Dis, 2018, 4, 467-477.

91. Makin, L.; Gluenz, E., cAMP signalling in trypanosomatids: Role in pathogenesis and as a

drug target. Trends Parasitol, 2015, 31, 373-379.

92. Alvarez-Rueda, N.; Biron, M.; Le Pape, P., Infectivity of Leishmania mexicana is associated

with differential expression of protein kinase C-like triggered during a cell-cell contact. PLoS One,

2009, 4, e7581.

93. Varela, M. R.; Ochoa, R.; Muskus, C. E.; Muro, A.; Mollinedo, F., Identification of a

RAC/AKT-like gene in Leishmania parasites as a putative therapeutic target in leishmaniasis.

Parasite Vectors, 2017, 10, 458.

94. Tirado-Duarte, D.; Marin-Villa, M.; Ochoa, R.; Blandon-Fuentes, G.; Soares, M. J.; Robledo,

S. M.; Varela-Miranda, R. E., The Akt-like kinase of Leishmania panamensis: As a new molecular

target for drug discovery. Acta Trop, 2018, 177, 171-178.

95. Ma, J.; Benz, C.; Grimaldi, R.; Stockdale, C.; Wyatt, P.; Frearson, J.; Hammarton, T. C.,

Nuclear DBF-2-related kinases are essential regulators of cytokinesis in bloodstream stage

Trypanosoma brucei. J Biol Chem, 2010, 285, 15356-15368.

96. Mora, A.; Komander, D.; van Aalten, D. M.; Alessi, D. R., PDK1, the master regulator of

AGC kinase signal transduction. Semin Cell Dev Biol, 2004, 15, 161-170.

97. Dittrich, A. C.; Devarenne, T. P., Perspectives in PDK1 evolution: Insights from

photosynthetic and non-photosynthetic organisms. Plant Signal Behav, 2012, 7, 642-649.

98. Digirolamo, F. A.; Miranda, M. R.; Bouvier, L. A.; Camara, M. M.; Canepa, G. E.; Pereira, C.

A., The mammalian TOR pathway is present in Trypanosoma cruzi. In silico reconstruction and

possible functions. Medicina (B Aires), 2012, 72, 221-226.

Page 167: Protein-kinase inhibitors as potential leishmanicidal drugs

152

99. Ramakrishnan, S.; Docampo, R., Membrane proteins in trypanosomatids involved in Ca(2+)

homeostasis and signaling. Genes (Basel), 2018, 9, 304.

100. Kato, K.; Sugi, T.; Takemae, H.; Takano, R.; Gong, H.; Ishiwa, A.; Horimoto, T.; Akashi, H.,

Characterization of a Toxoplasma gondii calcium calmodulin-dependent protein kinase homolog.

Parasite Vectors, 2016, 9, 405.

101. Andresen, C.; Niklasson, M.; Cassman Eklof, S.; Wallner, B.; Lundstrom, P., Biophysical

characterization of the calmodulin-like domain of Plasmodium falciparum calcium dependent

protein kinase 3. PLoS One, 2017, 12, e0181721.

102. Szoor, B., Trypanosomatid protein phosphatases. Mol Biochem Parasitol, 2010, 173, 53-63.

103. Dichiara, M.; Marrazzo, A.; Prezzavento, O.; Collina, S.; Rescifina, A.; Amata, E.,

Repurposing of human kinase inhibitors in neglected protozoan diseases. ChemMedChem, 2017, 12,

1235-1253.

104. Wyllie, S.; Thomas, M.; Patterson, S.; Crouch, S.; De Rycker, M.; Lowe, R.; Gresham, S.;

Urbaniak, M. D.; Otto, T. D.; Stojanovski, L.; Simeons, F. R. C.; Manthri, S.; MacLean, L. M.;

Zuccotto, F.; Homeyer, N.; Pflaumer, H.; Boesche, M.; Sastry, L.; Connolly, P.; Albrecht, S.;

Berriman, M.; Drewes, G.; Gray, D. W.; Ghidelli-Disse, S.; Dixon, S.; Fiandor, J. M.; Wyatt, P. G.;

Ferguson, M. A. J.; Fairlamb, A. H.; Miles, T. J.; Read, K. D.; Gilbert, I. H., Cyclin-dependent kinase

12 is a drug target for visceral leishmaniasis. Nature, 2018, 560, 192-197.

105. Naula, C.; Parsons, M.; Mottram, J. C., Protein kinases as drug targets in trypanosomes and

Leishmania. Biochim Biophys Acta, 2005, 1754, 151-159.

106. Saravanan, P.; Venkatesan, S. K.; Mohan, C. G.; Patra, S.; Dubey, V. K., Mitogen-activated

protein kinase 4 of Leishmania parasite as a therapeutic target. Eur J Med Chem, 2010, 45, 5662-

5670.

107. Raj, S.; Saha, G.; Sasidharan, S.; Dubey, V. K.; Saudagar, P., Biochemical characterization

and chemical validation of Leishmania MAP Kinase-3 as a potential drug target. Sci Rep, 2019, 9,

16209.

108. Kaur, P.; Garg, M.; Hombach-Barrigah, A.; Clos, J.; Goyal, N., MAPK1 of Leishmania

donovani interacts and phosphorylates HSP70 and HSP90 subunits of foldosome complex. Sci Rep,

2017, 7, 10202.

Page 168: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

153

109. Wang, Q.; Melzer, I. M.; Kruse, M.; Sander-Juelch, C.; Wiese, M., LmxMPK4, a mitogen-

activated protein (MAP) kinase homologue essential for promastigotes and amastigotes of

Leishmania mexicana. Kinetoplastid Biol Dis, 2005, 4, 6.

110. Maurer, U.; Preiss, F.; Brauns-Schubert, P.; Schlicher, L.; Charvet, C., GSK-3 - at the

crossroads of cell death and survival. J Cell Sci, 2014, 127, 1369-1378.

111. Xingi, E.; Smirlis, D.; Myrianthopoulos, V.; Magiatis, P.; Grant, K. M.; Meijer, L.; Mikros,

E.; Skaltsounis, A. L.; Soteriadou, K., 6-Br-5methylindirubin-3'oxime (5-Me-6-BIO) targeting the

leishmanial Glycogen Synthase Kinase-3 (GSK-3) short form affects cell-cycle progression and

induces apoptosis-like death: Exploitation of GSK-3 for treating leishmaniasis. Int J Parasitol, 2009,

39, 1289-1303.

112. Ghosh, G.; Adams, J. A., Phosphorylation mechanism and structure of serine-arginine protein

kinases. FEBS J, 2011, 278, 587-597.

113. Hombach-Barrigah, A.; Bartsch, K.; Smirlis, D.; Rosenqvist, H.; MacDonald, A.; Dingli, F.;

Loew, D.; Spath, G. F.; Rachidi, N.; Wiese, M.; Clos, J., Leishmania donovani 90 kD Heat Shock

Protein - Impact of phosphosites on parasite fitness, infectivity and Casein Kinase affinity. Sci Rep,

2019, 9, 5074.

114. Zylbersztejn, A. M.; de Morais, C. G.; Lima, A. K.; Souza, J. E.; Lopes, A. H.; Da-Silva, S.

A.; Silva-Neto, M. A.; Dutra, P. M., CK2 secreted by Leishmania braziliensis mediates macrophage

association invasion: A comparative study between virulent and avirulent promastigotes. Biomed Res

Int, 2015, 2015, 167323.

115. Broekhuis, J. R.; Verhey, K. J.; Jansen, G., Regulation of cilium length and intraflagellar

transport by the RCK-kinases ICK and MOK in renal epithelial cells. PLoS One, 2014, 9, e108470.

116. Jiang, Y. Y.; Maier, W.; Baumeister, R.; Minevich, G.; Joachimiak, E.; Wloga, D.; Ruan, Z.;

Kannan, N.; Bocarro, S.; Bahraini, A.; Vasudevan, K. K.; Lechtreck, K.; Orias, E.; Gaertig, J.,

LF4/MOK and a CDK-related kinase regulate the number and length of cilia in Tetrahymena. PLoS

Genet, 2019, 15, e1008099.

117. Wang, Y.; Ren, Y.; Pan, J., Regulation of flagellar assembly and length in Chlamydomonas

by LF4, a MAPK-related kinase. FASEB J, 2019, 33, 6431-6441.

118. Rachidi, N.; Taly, J. F.; Durieu, E.; Leclercq, O.; Aulner, N.; Prina, E.; Pescher, P.; Notredame,

C.; Meijer, L.; Spath, G. F., Pharmacological assessment defines Leishmania donovani Casein

Page 169: Protein-kinase inhibitors as potential leishmanicidal drugs

154

Kinase 1 as a drug target and reveals important functions in parasite viability and intracellular

infection. Antimicrob Agents Chemother, 2014, 58, 1501-1515.

119. Dan-Goor, M.; Nasereddin, A.; Jaber, H.; Jaffe, C. L., Identification of a secreted Casein

Kinase 1 in Leishmania donovani: Effect of protein over expression on parasite growth and virulence.

PLoS One, 2013, 8, e79287.

120. Ikezu, S.; Ikezu, T., Tau-tubulin kinase. Front Mol Neurosci, 2014, 7, 33.

121. Brumlik, M. J.; Pandeswara, S.; Ludwig, S. M.; Murthy, K.; Curiel, T. J., Parasite mitogen-

activated protein kinases as drug discovery targets to treat human protozoan pathogens. J Signal

Transduct, 2011, 2011, 971968.

122. Jones, N. G.; Thomas, E. B.; Brown, E.; Dickens, N. J.; Hammarton, T. C.; Mottram, J. C.,

Regulators of Trypanosoma brucei cell cycle progression and differentiation identified using a

kinome-wide RNAi screen. PLoS Pathog, 2014, 10, e1003886.

123. Chhajer, R.; Bhattacharyya, A.; Didwania, N.; Shadab, M.; Das, N.; Palit, P.; Vaidya, T.; Ali,

N., Leishmania donovani Aurora kinase: A promising therapeutic target against visceral

leishmaniasis. Biochim Biophys Acta, 2016, 1860, 1973-1988.

124. Hammarton, T. C.; Kramer, S.; Tetley, L.; Boshart, M.; Mottram, J. C., Trypanosoma brucei

Polo-like kinase is essential for basal body duplication, kDNA segregation and cytokinesis. Mol

Microbiol, 2007, 65, 1229-1248.

125. Boynak, N. Y.; Rojas, F.; D'Alessio, C.; Vilchez Larrea, S. C.; Rodriguez, V.; Ghiringhelli, P.

D.; Tellez-Inon, M. T., Identification of a Wee1-like kinase gene essential for procyclic

Trypanosoma brucei survival. PLoS One, 2013, 8, e79364.

126. Vijayakumar, S.; Das, P., Recent progress in drug targets and inhibitors towards combating

leishmaniasis. Acta Trop, 2018, 181, 95-104.

127. Sundar, S.; Singh, A., Chemotherapeutics of visceral leishmaniasis: Present and future

developments. Parasitology, 2018, 145, 481-489.

128. Hendrickx, S.; Caljon, G.; Maes, L., Need for sustainable approaches in antileishmanial drug

discovery. Parasitol Res, 2019, 118, 2743-2752.

129. Kapil, S.; Singh, P. K.; Silakari, O., An update on small molecule strategies targeting

leishmaniasis. Eur J Med Chem, 2018, 157, 339-367.

Page 170: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

155

130. Haldar, A. K.; Sen, P.; Roy, S., Use of antimony in the treatment of leishmaniasis: Current

status and future directions. Mol Biol Int, 2011, 2011, 571242.

131. Kip, A. E.; Schellens, J. H. M.; Beijnen, J. H.; Dorlo, T. P. C., Clinical pharmacokinetics of

systemically administered antileishmanial drugs. Clin Pharmacokinet, 2018, 57, 151-176.

132. Kumar, A.; Pandey, S. C.; Samant, M., Slow pace of antileishmanial drug development.

Parasitology Open, 2018, 4, 1-11.

133. Vincent, I. M.; Weidt, S.; Rivas, L.; Burgess, K.; Smith, T. K.; Ouellette, M., Untargeted

metabolomic analysis of miltefosine action in Leishmania infantum reveals changes to the internal

lipid metabolism. Int J Parasitol Drugs Drug Resist, 2014, 4, 20-27.

134. Field, M. C.; Horn, D.; Fairlamb, A. H.; Ferguson, M. A. J.; Gray, D. W.; Read, K. D.; De

Rycker, M.; Torrie, L. S.; Wyatt, P. G.; Wyllie, S.; Gilbert, I. H., Anti-trypanosomatid drug

discovery: An ongoing challenge and a continuing need. Nat Rev Microbiol, 2017, 15, 217-231.

135. Charlton, R. L.; Rossi-Bergmann, B.; Denny, P. W.; Steel, P. G., Repurposing as a strategy

for the discovery of new anti-leishmanials: the-state-of-the-art. Parasitology, 2018, 145, 219-236.

136. Rao, S. P. S.; Barrett, M. P.; Dranoff, G.; Faraday, C. J.; Gimpelewicz, C. R.; Hailu, A.; Jones,

C. L.; Kelly, J. M.; Lazdins-Helds, J. K.; Maser, P.; Mengel, J.; Mottram, J. C.; Mowbray, C. E.;

Sacks, D. L.; Scott, P.; Spath, G. F.; Tarleton, R. L.; Spector, J. M.; Diagana, T. T., Drug discovery

for kinetoplastid diseases: Future directions. ACS Infect Dis, 2019, 5, 152-157.

137. Mowbray, C., Anti-leishmanial drug discovery: Past, present and future perspectives, in Drug

discovery for leishmaniasis, ed. L. Rivas; C. Gil, Royal Society of Chemistry, United Kingdom,

2017, pp. 24-36.

138. Sebastian-Perez, V.; Roca, C.; Awale, M.; Reymond, J. L.; Martinez, A.; Gil, C.; Campillo,

N. E., Medicinal and Biological Chemistry (MBC) library: An efficient source of new hits. J Chem

Inf Model, 2017, 57, 2143-2151.

139. Luque-Ortega, J. R.; Rivero-Lezcano, O. M.; Croft, S. L.; Rivas, L., In vivo monitoring of

intracellular ATP levels in Leishmania donovani promastigotes as a rapid method to screen drugs

targeting bioenergetic metabolism. Antimicrob Agents Chemother, 2001, 45, 1121-1125.

140. Ojo, K. K.; Arakaki, T. L.; Napuli, A. J.; Inampudi, K. K.; Keyloun, K. R.; Zhang, L.; Hol, W.

G.; Verlinde, C. L.; Merritt, E. A.; Van Voorhis, W. C., Structure determination of Glycogen

Page 171: Protein-kinase inhibitors as potential leishmanicidal drugs

156

Synthase Kinase-3 from Leishmania major and comparative inhibitor structure-activity relationships

with Trypanosoma brucei GSK-3. Mol Biochem Parasitol, 2011, 176, 98-108.

141. Luque-Ortega, J. R.; Rivas, L., Characterization of the leishmanicidal activity of antimicrobial

peptides in Antimicrobial peptides, ed. A. Giuliani, A. C. Rinaldi, Humana Press, USA, 2010, 618,

393-420.

142. Drugs for Neglected Diseases initiative. DNDi R&D portfolio June 2019

https://www.dndi.org/diseases-projects/portfolio/. Accessed September 2020.

143. Patel, P.; Woodgett, J. R., Glycogen synthase kinase 3: A kinase for all pathways? in Protein

kinases in development and disease, ed. A. Jenny, Elsevier, Netherlands, 2017, pp. 277-302.

144. Noori, M. S.; Bhatt, P. M.; Courreges, M. C.; Ghazanfari, D.; Cuckler, C.; Orac, C. M.;

McMills, M. C.; Schwartz, F. L.; Deosarkar, S. P.; Bergmeier, S. C.; McCall, K. D.; Goetz, D. J.,

Identification of a novel selective and potent inhibitor of Glycogen Synthase Kinase-3. Am J Physiol

Cell Physiol, 2019, 317, C1289-C1303.

145. Efstathiou, A.; Gaboriaud-Kolar, N.; Myrianthopoulos, V.; Vougogiannopoulou, K.; Subota,

I.; Aicher, S.; Mikros, E.; Bastin, P.; Skaltsounis, A. L.; Soteriadou, K.; Smirlis, D., Indirubin

analogues inhibit Trypanosoma brucei Glycogen Synthase Kinase 3 short and T. brucei growth.

Antimicrob Agents Chemother, 2019, 63, e02065-02018.

146. Efstathiou, A.; Gaboriaud-Kolar, N.; Smirlis, D.; Myrianthopoulos, V.; Vougogiannopoulou,

K.; Alexandratos, A.; Kritsanida, M.; Mikros, E.; Soteriadou, K.; Skaltsounis, A. L., An inhibitor-

driven study for enhancing the selectivity of indirubin derivatives towards leishmanial Glycogen

Synthase Kinase-3 over leishmanial cdc2-related protein kinase 3. Parasite Vectors, 2014, 7, 234.

147. Osolodkin, D. I.; Zakharevich, N. V.; Palyulin, V. A.; Danilenko, V. N.; Zefirov, N. S.,

Bioinformatic analysis of Glycogen Synthase Kinase 3: Human versus parasite kinases.

Parasitology, 2011, 138, 725-735.

148. Garcia, I.; Fall, Y.; Gomez, G.; Gonzalez-Diaz, H., First computational chemistry multi-target

model for anti-Alzheimer, anti-parasitic, anti-fungi, and anti-bacterial activity of GSK-3 inhibitors

in vitro, in vivo, and in different cellular lines. Mol Divers, 2011, 15, 561-567.

149. Rosano, G. L.; Ceccarelli, E. A., Recombinant protein expression in Escherichia coli:

Advances and challenges. Front Microbiol, 2014, 5, 172.

Page 172: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

157

150. John von Freyend, S.; Rosenqvist, H.; Fink, A.; Melzer, I. M.; Clos, J.; Jensen, O. N.; Wiese,

M., LmxMPK4, an essential mitogen-activated protein kinase of Leishmania mexicana is

phosphorylated and activated by the STE7-like protein kinase LmxMKK5. Int J Parasitol, 2010, 40,

969-978.

151. Grant, K. M.; Dunion, M. H.; Yardley, V.; Skaltsounis, A. L.; Marko, D.; Eisenbrand, G.;

Croft, S. L.; Meijer, L.; Mottram, J. C., Inhibitors of Leishmania mexicana CRK3 cyclin-dependent

kinase: Chemical library screen and antileishmanial activity. Antimicrob Agents Chemother, 2004,

48, 3033-3042.

152. Cann, M. L.; McDonald, I. M.; East, M. P.; Johnson, G. L.; Graves, L. M., Measuring kinase

activity-A global challenge. J Cell Biochem, 2017, 118, 3595-3606.

153. Baki, A.; Bielik, A.; Molnar, L.; Szendrei, G.; Keseru, G. M., A high throughput luminescent

assay for Glycogen Synthase Kinase-3beta inhibitors. Assay Drug Dev Technol, 2007, 5, 75-83.

154. Ojo, K. K.; Larson, E. T.; Keyloun, K. R.; Castaneda, L. J.; Derocher, A. E.; Inampudi, K. K.;

Kim, J. E.; Arakaki, T. L.; Murphy, R. C.; Zhang, L.; Napuli, A. J.; Maly, D. J.; Verlinde, C. L.;

Buckner, F. S.; Parsons, M.; Hol, W. G.; Merritt, E. A.; Van Voorhis, W. C., Toxoplasma gondii

Calcium-dependent Protein Kinase 1 is a target for selective kinase inhibitors. Nat Struct Mol Biol,

2010, 17, 602-607.

155. Benek, O.; Hroch, L.; Aitken, L.; Gunn-Moore, F.; Vinklarova, L.; Kuca, K.; Perez, D. I.;

Perez, C.; Martinez, A.; Fisar, Z.; Musilek, K., 1-(Benzo[d]thiazol-2-yl)-3-phenylureas as dual

inhibitors of Casein Kinase 1 and ABAD enzymes for treatment of neurodegenerative disorders. J

Enzyme Inhib Med Chem, 2018, 33, 665-670.

156. Sutherland, C.; Leighton, I. A.; Cohen, P., Inactivation of Glycogen Synthase Kinase-3 beta

by phosphorylation: New kinase connections in insulin and growth-factor signalling. Biochem J,

1993, 296, 15-19.

157. Droucheau, E.; Primot, A.; Thomas, V.; Mattei, D.; Knockaert, M.; Richardson, C.;

Sallicandro, P.; Alano, P.; Jafarshad, A.; Baratte, B.; Kunick, C.; Parzy, D.; Pearl, L.; Doerig, C.;

Meijer, L., Plasmodium falciparum Glycogen Synthase Kinase-3: Molecular model, expression,

intracellular localisation and selective inhibitors. Biochim Biophys Acta, 2004, 1697, 181-196.

158. Estañ, N. Búsqueda de inhibidores de quinasas como posibles fármacos para el tratamiento

de la leishmaniasis. Master thesis, Universidad de Alcala de Henares, Spain, 2017.

Page 173: Protein-kinase inhibitors as potential leishmanicidal drugs

158

159. Sebastian-Perez, V. Drug discovery for infectious diseases: Target-based and ligand-based

approaches. PhD thesis, Universidad Complutense de Madrid, Spain, 2019.

160. Blum, J. J., Energy metabolism in Leishmania. J Bioenerg Biomembr, 1994, 26, 147-155.

161. Saar, Y.; Ransford, A.; Waldman, E.; Mazareb, S.; Amin-Spector, S.; Plumblee, J.; Turco, S.

J.; Zilberstein, D., Characterization of developmentally-regulated activities in axenic amastigotes of

Leishmania donovani. Mol Biochem Parasitol, 1998, 95, 9-20.

162. Bahrami, S.; Hatam, G. R.; Razavi, M.; Nazifi, S., In vitro cultivation of axenic amastigotes

and the comparison of antioxidant enzymes at different stages of Leishmania tropica. Trop Biomed,

2011, 28, 411-417.

163. Pan, A. A.; Duboise, S. M.; Eperon, S.; Rivas, L.; Hodgkinson, V.; Traub-Cseko, Y.;

McMahon-Pratt, D., Developmental life cycle of Leishmania--Cultivation and characterization of

cultured extracellular amastigotes. J Eukaryot Microbiol, 1993, 40, 213-223.

164. Chanmol, W.; Jariyapan, N.; Somboon, P.; Bates, M. D.; Bates, P. A., Axenic amastigote

cultivation and in vitro development of Leishmania orientalis. Parasitol Res, 2019, 118, 1885-1897.

165. Vermeersch, M.; da Luz, R. I.; Tote, K.; Timmermans, J. P.; Cos, P.; Maes, L., In vitro

susceptibilities of Leishmania donovani promastigote and amastigote stages to antileishmanial

reference drugs: Practical relevance of stage-specific differences. Antimicrob Agents Chemother,

2009, 53, 3855-3859.

166. Palic, S.; Bhairosing, P.; Beijnen, J. H.; Dorlo, T. P. C., Systematic review of host-mediated

activity of miltefosine in leishmaniasis through immunomodulation. Antimicrob Agents Chemother,

2019, 63, e02507-02518.

167. Martinez, A.; Alonso, M.; Castro, A.; Perez, C.; Moreno, F. J., First non-ATP competitive

Glycogen Synthase Kinase 3 beta (GSK-3beta) inhibitors: Thiadiazolidinones (TDZD) as potential

drugs for the treatment of Alzheimer's disease. J Med Chem, 2002, 45, 1292-1299.

168. Palomo, V.; Perez, D. I.; Perez, C.; Morales-Garcia, J. A.; Soteras, I.; Alonso-Gil, S.; Encinas,

A.; Castro, A.; Campillo, N. E.; Perez-Castillo, A.; Gil, C.; Martinez, A., 5-imino-1,2,4-thiadiazoles:

First small molecules as substrate competitive inhibitors of Glycogen Synthase Kinase 3. J Med

Chem, 2012, 55, 1645-1661.

Page 174: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

159

169. Palomo, V.; Soteras, I.; Perez, D. I.; Perez, C.; Gil, C.; Campillo, N. E.; Martinez, A.,

Exploring the binding sites of Glycogen Synthase Kinase 3. Identification and characterization of

allosteric modulation cavities. J Med Chem, 2011, 54, 8461-8470.

170. Conde, S.; Perez, D. I.; Martinez, A.; Perez, C.; Moreno, F. J., Thienyl and phenyl alpha-

halomethyl ketones: New inhibitors of Glycogen Synthase Kinase (GSK-3beta) from a library of

compound searching. J Med Chem, 2003, 46, 4631-4633.

171. Perez, D. I.; Palomo, V.; Perez, C.; Gil, C.; Dans, P. D.; Luque, F. J.; Conde, S.; Martinez, A.,

Switching reversibility to irreversibility in Glycogen Synthase Kinase 3 inhibitors: Clues for specific

design of new compounds. J Med Chem, 2011, 54, 4042-4056.

172. Perez-Domper, P.; Palomo, V.; Gradari, S.; Gil, C.; de Ceballos, M. L.; Martinez, A.; Trejo,

J. L., The GSK-3-inhibitor VP2.51 produces antidepressant effects associated with adult

hippocampal neurogenesis. Neuropharmacology, 2017, 116, 174-187.

173. Salado, I. G.; Redondo, M.; Bello, M. L.; Perez, C.; Liachko, N. F.; Kraemer, B. C.; Miguel,

L.; Lecourtois, M.; Gil, C.; Martinez, A.; Perez, D. I., Protein kinase CK-1 inhibitors as new potential

drugs for amyotrophic lateral sclerosis. J Med Chem, 2014, 57, 2755-2772.

174. Salado, I. G. Nuevos inhibidores de proteína quinasa como tratamiento innovador para

enfermedades neurodegenerativas. PhD thesis, Universidad Complutense de Madrid, Spain, 2014.

175. Salado, I. G.; Zaldivar-Diez, J.; Sebastian-Perez, V.; Li, L.; Geiger, L.; Gonzalez, S.;

Campillo, N. E.; Gil, C.; Morales, A. V.; Perez, D. I.; Martinez, A., Leucine rich repeat kinase 2

(LRRK2) inhibitors based on indolinone scaffold: Potential pro-neurogenic agents. Eur J Med Chem,

2017, 138, 328-342.

176. Dominguez, J. M.; Fuertes, A.; Orozco, L.; del Monte-Millan, M.; Delgado, E.; Medina, M.,

Evidence for irreversible inhibition of Glycogen Synthase Kinase-3beta by tideglusib. J Biol Chem,

2012, 287, 893-904.

177. Martinez de Iturrate, P.; Sebastian-Perez, V.; Nacher-Vazquez, M.; Tremper, C. S.; Smirlis,

D.; Martin, J.; Martinez, A.; Campillo, N. E.; Rivas, L.; Gil, C., Towards discovery of new

leishmanicidal scaffolds able to inhibit Leishmania GSK-3. J Enzyme Inhib Med Chem, 2020, 35,

199-210.

178. Datta, R.; Das, I.; Sen, B.; Chakraborty, A.; Adak, S.; Mandal, C.; Datta, A. K., Mutational

analysis of the active-site residues crucial for catalytic activity of adenosine kinase from Leishmania

donovani. Biochem J, 2005, 387, 591-600.

Page 175: Protein-kinase inhibitors as potential leishmanicidal drugs

160

179. Vermelho, A. B.; Capaci, G. R.; Rodrigues, I. A.; Cardoso, V. S.; Mazotto, A. M.; Supuran,

C. T., Carbonic anhydrases from Trypanosoma and Leishmania as anti-protozoan drug targets.

Bioorg Med Chem, 2017, 25, 1543-1555.

180. Frezard, F.; Monte-Neto, R.; Reis, P. G., Antimony transport mechanisms in resistant

Leishmania parasites. Biophys Rev, 2014, 6, 119-132.

181. Senger, M. R.; Fraga, C. A.; Dantas, R. F.; Silva, F. P., Jr., Filtering promiscuous compounds

in early drug discovery: Is it a good idea? Drug Discov Today, 2016, 21, 868-872.

182. Vasudevan, A.; Argiriadi, M. A.; Baranczak, A.; Friedman, M. M.; Gavrilyuk, J.; Hobson, A.

D.; Hulce, J. J.; Osman, S.; Wilson, N. S., Covalent binders in drug discovery. Prog Med Chem,

2019, 58, 1-62.

183. LoGiudice, N.; Le, L.; Abuan, I.; Leizorek, Y.; Roberts, S. C., Alpha-difluoromethylornithine,

an irreversible inhibitor of polyamine biosynthesis, as a therapeutic strategy against

hyperproliferative and infectious diseases. Med Sci (Basel), 2018, 6, 12.

184. Hongisto, V.; Vainio, J. C.; Thompson, R.; Courtney, M. J.; Coffey, E. T., The Wnt pool of

Glycogen Synthase Kinase 3beta is critical for trophic-deprivation-induced neuronal death. Mol Cell

Biol, 2008, 28, 1515-1527.

185. Doyle, P. S.; Engel, J. C.; Gam, A. A.; Dvorak, J. A., Leishmania mexicana mexicana:

Quantitative analysis of the intracellular cycle. Parasitology, 1989, 99 Pt 3, 311-316.

186. Marchena, M.; Villarejo-Zori, B.; Zaldivar-Diez, J.; Palomo, V.; Gil, C.; Hernandez-Sanchez,

C.; Martinez, A.; de la Rosa, E. J., Small molecules targeting Glycogen Synthase Kinase 3 as

potential drug candidates for the treatment of retinitis pigmentosa. J Enzyme Inhib Med Chem, 2017,

32, 522-526.

187. Mathuram, T. L.; Reece, L. M.; Cherian, K. M., GSK-3 inhibitors: A double-edged sword? -

An update on tideglusib. Drug Res (Stuttg), 2018, 68, 436-443.

188. Schneider, G.; Bohm, H. J., Virtual screening and fast automated docking methods. Drug

Discov Today, 2002, 7, 64-70.

189. Torrie, L. S.; Robinson, D. A.; Thomas, M. G.; Hobrath, J. V.; Shepherd, S. M.; Post, J. M.;

Ko, E. J.; Ferreira, R. A.; Mackenzie, C. J.; Wrobel, K.; Edwards, D. P.; Gilbert, I. H.; Gray, D. W.;

Fairlamb, A. H.; De Rycker, M., Discovery of an allosteric binding site in kinetoplastid methionyl-

tRNA synthetase. ACS Infect Dis, 2020, 6, 1044-1057.

Page 176: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

161

190. Sundar, S.; Singh, B., Emerging therapeutic targets for treatment of leishmaniasis. Expert Opin

Ther Targets, 2018, 22, 467-486.

191. Ortiz, D.; Forquer, I.; Boitz, J.; Soysa, R.; Elya, C.; Fulwiler, A.; Nilsen, A.; Polley, T.; Riscoe,

M. K.; Ullman, B.; Landfear, S. M., Targeting the cytochrome bc1 complex of Leishmania parasites

for discovery of novel drugs. Antimicrob Agents Chemother, 2016, 60, 4972-4982.

192. Pinto, E. G.; Santos, I. O.; Schmidt, T. J.; Borborema, S. E.; Ferreira, V. F.; Rocha, D. R.;

Tempone, A. G., Potential of 2-hydroxy-3-phenylsulfanylmethyl-[1,4]-naphthoquinones against

Leishmania (L.) infantum: Biological activity and structure-activity relationships. PLoS One, 2014,

9, e105127.

193. Raj, S.; Sasidharan, S.; Dubey, V. K.; Saudagar, P., Identification of lead molecules against

potential drug target protein MAPK4 from L. donovani: An in-silico approach using docking,

molecular dynamics and binding free energy calculation. PLoS One, 2019, 14, e0221331.

194. Hanson, S. M.; Georghiou, G.; Thakur, M. K.; Miller, W. T.; Rest, J. S.; Chodera, J. D.;

Seeliger, M. A., What makes a kinase promiscuous for inhibitors? Cell Chem Biol, 2019, 26, 390-

399 e395.

195. Cavazzuti, A.; Paglietti, G.; Hunter, W. N.; Gamarro, F.; Piras, S.; Loriga, M.; Allecca, S.;

Corona, P.; McLuskey, K.; Tulloch, L.; Gibellini, F.; Ferrari, S.; Costi, M. P., Discovery of potent

pteridine reductase inhibitors to guide antiparasite drug development. Proc Natl Acad Sci U S A,

2008, 105, 1448-1453.

196. Kunnimalaiyaan, S.; Schwartz, V. K.; Jackson, I. A.; Clark Gamblin, T.; Kunnimalaiyaan, M.,

Antiproliferative and apoptotic effect of LY2090314, a GSK-3 inhibitor, in neuroblastoma in vitro.

BMC Cancer, 2018, 18, 560.

197. Ugolkov, A. V.; Bondarenko, G. I.; Dubrovskyi, O.; Berbegall, A. P.; Navarro, S.; Noguera,

R.; O'Halloran, T. V.; Hendrix, M. J.; Giles, F. J.; Mazar, A. P., 9-ING-41, a small-molecule

Glycogen Synthase Kinase-3 inhibitor, is active in neuroblastoma. Anticancer Drugs, 2018, 29, 717-

724.

198. Yardley, V.; Koniordou, M., Drug assay methodology in leishmaniasis: From the microplate

to image analysis, in Drug discovery for leishmaniasis, ed. L. Rivas; C. Gil, Royal Society of

Chemistry, United Kingdom, 2017, pp. 57-76.

199. Seifert, K.; Escobar, P.; Croft, S. L., In vitro activity of anti-leishmanial drugs against

Leishmania donovani is host cell dependent. J Antimicrob Chemother, 2010, 65, 508-511.

Page 177: Protein-kinase inhibitors as potential leishmanicidal drugs

162

200. Hendrickx, S.; Van Bockstal, L.; Caljon, G.; Maes, L., In-depth comparison of cell-based

methodological approaches to determine drug susceptibility of visceral Leishmania isolates. PLoS

Negl Trop Dis, 2019, 13, e0007885.

201. Morimoto, A.; Uchida, K.; Chambers, J. K.; Sato, K.; Hong, J.; Sanjoba, C.; Matsumoto, Y.;

Yamagishi, J.; Goto, Y., Hemophagocytosis induced by Leishmania donovani infection is beneficial

to parasite survival within macrophages. PLoS Negl Trop Dis, 2019, 13, e0007816.

202. Sun, Y.; Chen, D.; Pan, Y.; Qu, W.; Hao, H.; Wang, X.; Liu, Z.; Xie, S., Nanoparticles for

antiparasitic drug delivery. Drug Deliv, 2019, 26, 1206-1221.

203. Bruni, N.; Stella, B.; Giraudo, L.; Della Pepa, C.; Gastaldi, D.; Dosio, F., Nanostructured

delivery systems with improved leishmanicidal activity: A critical review. Int J Nanomedicine, 2017,

12, 5289-5311.

204. Li, Y. H.; Wang, P. P.; Li, X. X.; Yu, C. Y.; Yang, H.; Zhou, J.; Xue, W. W.; Tan, J.; Zhu, F.,

The human kinome targeted by FDA approved multi-target drugs and combination products: A

comparative study from the drug-target interaction network perspective. PLoS One, 2016, 11,

e0165737.

205. Pieretti, S.; Haanstra, J. R.; Mazet, M.; Perozzo, R.; Bergamini, C.; Prati, F.; Fato, R.; Lenaz,

G.; Capranico, G.; Brun, R.; Bakker, B. M.; Michels, P. A.; Scapozza, L.; Bolognesi, M. L.; Cavalli,

A., Naphthoquinone derivatives exert their antitrypanosomal activity via a multi-target mechanism.

PLoS Negl Trop Dis, 2013, 7, e2012.

206. Braga, S. S., Multi-target drugs active against leishmaniasis: A paradigm of drug repurposing.

Eur J Med Chem, 2019, 183, 111660.

207. Talevi, A., Multi-target pharmacology: Possibilities and limitations of the "skeleton key

approach" from a medicinal chemist perspective. Front Pharmacol, 2015, 6, 205.

208. Dorlo, T. P.; Balasegaram, M.; Beijnen, J. H.; de Vries, P. J., Miltefosine: A review of its

pharmacology and therapeutic efficacy in the treatment of leishmaniasis. J Antimicrob Chemother,

2012, 67, 2576-2597.

209. Souder, D. C.; Anderson, R. M., An expanding GSK3 network: Implications for aging

research. Geroscience, 2019, 41, 369-382.

Page 178: Protein-kinase inhibitors as potential leishmanicidal drugs

Bibliography

163

210. Yang, K.; Chen, Z.; Gao, J.; Shi, W.; Li, L.; Jiang, S.; Hu, H.; Liu, Z.; Xu, D.; Wu, L., The

key roles of GSK-3beta in regulating mitochondrial activity. Cell Physiol Biochem, 2017, 44, 1445-

1459.

211. Benaim, G.; Paniz-Mondolfi, A. E.; Sordillo, E. M.; Martinez-Sotillo, N., Disruption of

intracellular calcium homeostasis as a therapeutic target against Trypanosoma cruzi. Front Cell Infect

Microbiol, 2020, 10, 46.

212. Alvarez-Fortes, E.; Ruiz-Perez, L. M.; Bouillaud, F.; Rial, E.; Rivas, L., Expression and

regulation of mitochondrial Uncoupling Protein 1 from brown adipose tissue in Leishmania major

promastigotes. Mol Biochem Parasitol, 1998, 93, 191-202.

213. Gerencser, A. A.; Chinopoulos, C.; Birket, M. J.; Jastroch, M.; Vitelli, C.; Nicholls, D. G.;

Brand, M. D., Quantitative measurement of mitochondrial membrane potential in cultured cells:

Calcium-induced de- and hyperpolarization of neuronal mitochondria. J Physiol, 2012, 590, 2845-

2871.

214. Alcazar-Fabra, M.; Navas, P.; Brea-Calvo, G., Coenzyme Q biosynthesis and its role in the

respiratory chain structure. Biochim Biophys Acta, 2016, 1857, 1073-1078.

215. Carvalho, L.; Luque-Ortega, J. R.; Lopez-Martin, C.; Castanys, S.; Rivas, L.; Gamarro, F.,

The 8-aminoquinoline analogue sitamaquine causes oxidative stress in Leishmania donovani

promastigotes by targeting succinate dehydrogenase. Antimicrob Agents Chemother, 2011, 55, 4204-

4210.

216. Carvalho, L.; Luque-Ortega, J. R.; Manzano, J. I.; Castanys, S.; Rivas, L.; Gamarro, F.,

Tafenoquine, an antiplasmodial 8-aminoquinoline, targets Leishmania respiratory complex III and

induces apoptosis. Antimicrob Agents Chemother, 2010, 54, 5344-5351.

217. Silva, A. M.; Oliveira, P. J., Evaluation of respiration with Clark-type electrode in isolated

mitochondria and permeabilized animal cells. Methods Mol Biol, 2018, 1782, 7-29.

218. Biocyclopedia. Oxygen electrodes

https://biocyclopedia.com/index/chem_lab_methods/oxygen_electrodes.php. Accessed September

2020.

219. Vercesi, A. E.; Bernardes, C. F.; Hoffmann, M. E.; Gadelha, F. R.; Docampo, R., Digitonin

permeabilization does not affect mitochondrial function and allows the determination of the

mitochondrial membrane potential of Trypanosoma cruzi in situ. J Biol Chem, 1991, 266, 14431-

14434.

Page 179: Protein-kinase inhibitors as potential leishmanicidal drugs

164

220. Monzote, L.; Stamberg, W.; Patel, A.; Rosenau, T.; Maes, L.; Cos, P.; Gille, L., Synthetic

chromanol derivatives and their interaction with complex III in mitochondria from bovine, yeast, and

Leishmania. Chem Res Toxicol, 2011, 24, 1678-1685.

221. Monzote, L.; Geroldinger, G.; Tonner, M.; Scull, R.; De Sarkar, S.; Bergmann, S.; Bacher, M.;

Staniek, K.; Chatterjee, M.; Rosenau, T.; Gille, L., Interaction of ascaridole, carvacrol, and

caryophyllene oxide from essential oil of Chenopodium ambrosioides L. with mitochondria in

Leishmania and other eukaryotes. Phytother Res, 2018, 32, 1729-1740.

222. Menna-Barreto, R. F. S., Cell death pathways in pathogenic trypanosomatids: Lessons of

(over)kill. Cell Death Dis, 2019, 10, 93.

223. Basmaciyan, L.; Azas, N.; Casanova, M., Different apoptosis pathways in Leishmania

parasites. Cell Death Discov, 2018, 4, 27.

224. Luque-Ortega, J. R.; Rivas, L., Miltefosine (hexadecylphosphocholine) inhibits cytochrome c

oxidase in Leishmania donovani promastigotes. Antimicrob Agents Chemother, 2007, 51, 1327-

1332.

225. Zuo, X.; Djordjevic, J. T.; Bijosono Oei, J.; Desmarini, D.; Schibeci, S. D.; Jolliffe, K. A.;

Sorrell, T. C., Miltefosine induces apoptosis-like cell death in yeast via Cox9p in cytochrome c

oxidase. Mol Pharmacol, 2011, 80, 476-485.

Page 180: Protein-kinase inhibitors as potential leishmanicidal drugs

RESULTS DISSEMINATION

Page 181: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 182: Protein-kinase inhibitors as potential leishmanicidal drugs

Results dissemination

167

Results dissemination

Publications from the PhD thesis

1. Martínez de Iturrate, P.; Sebastián-Perez, V.; Nácher-Vázquez, M.; Tremper, C. S.;

Smirlis, D.; Martín, J.; Martínez, A.; Campillo, N. E.; Rivas, L.; Gil, C., Towards discovery

of new leishmanicidal scaffolds able to inhibit Leishmania GSK-3. J Enzyme Inhib Med

Chem, 2020, 35, 199-210.

Participation in Conferences, Symposiums and Congresses

1. Poster at XXIV International Symposium on Medicinal Chemistry. Manchester, UK,

2016. “Drug repurposing of human kinase inhibitors as new hits against Leishmania”

(Abstract book, P155). Authors: Gil, C.; Abengózar, M. A.; Martínez de Iturrate, P.;

Sebastián-Pérez, V.; Martínez, A.; Campillo, N. E.; Rivas, L.

2. Poster at 3rd COST Action CM1307 Conference. Madrid, Spain, 2016. “Drug

repurposing of human kinase inhibitors as new hits against Leishmania” (Abstract book,

P.10.). Authors: Gil, C.; Abengózar, M. A.; Martínez de Iturrate, P.; Sebastián-Pérez, V.;

Martínez, A.; Campillo, N. E.; Rivas, L.

3. Poster at WorldLeish6. Toledo, Spain, 2017. “New small heterocyclic scaffolds with

leishmanicidal activity” (Abstract book, C1313). Authors: Rivas, L.; Martínez de Iturrate,

P.; Abengózar, M. A.; Sebastián-Pérez, V.; Nácher, M.; Martínez, A.; Gil, C.

4. Poster at EUROPIN Summer School on Drug Design. Vienna, Austria, 2017.

“Targeting protein kinases in Leishmania as a novel approach to discover new drugs”

(Abstract and certification). Authors: Sebastián-Pérez, V.; Estañ, N.; Nácher-Vázquez, M.;

Martínez de Iturrate, P.; Rivas, L.; Campillo, N.E.; Gil, C.

Page 183: Protein-kinase inhibitors as potential leishmanicidal drugs

168

5. Poster at I Jornada PhDay Facultad de Farmacia UCM. Madrid, Spain 2017. “Protein

kinase inhibitors as leads for new chemotherapeutics against Leishmania” (Abstract book,

P29). Authors: Martínez de Iturrate, P.; Sebastián-Pérez, V.; Nácher-Vázquez, M.;

Abengózar, M. A.; Martínez, A.; Campillo, N.E.; Rivas, L.; Gil, C.

6. Oral Presentation at I Jornada PhDay Facultad de Farmacia UCM. Madrid, Spain 2017.

“Protein kinase inhibitors as leads for new chemotherapeutics against Leishmania” (Abstract

book and certification).

7. Poster at 20th European Bioenergetics Conference. Budapest, Hungary, 2018. “The

energy metabolism of Leishmania as a drug target” (Abstract book, P10/P16). Authors: Rial,

E.; Lastra-Romero, A.; Martínez-de-Iturrate, P.; Nácher-Vázquez, M.; Rivas, L.

8. Poster at 41 Congreso de la Sociedad Española de Bioquímica y Biología Molecular.

Santander, Spain, 2018. “In house libraries of human protein kinase inhibitors as a source

for new leishmanicidal agents” (Abstract book, P0232). Authors: Rivas, L.; Martínez de

Iturrate, P.; Sebastián-Pérez, V.; Nácher-Vázquez, M.; Tremper, C.; Lastra, A.; Rial, E.;

Campillo, N.E.; Gil, C.

9. Poster at 6th Young Reasearchers Symposium. Madrid, Spain, 2019. “The quest for

new Leishmania GSK3 inhibitors from large drug libraries: Leishbox as a representative

example” (Abstract book, P14). Authors: Tremper, C.S.; Sebastián-Pérez, V.; Martínez de

Iturrate, P.; Nácher-Vázquez, M.; Campillo, N.E.; Gil, C.; Rivas, L.

10. Poster at 6th Young Researchers Symposium. Madrid, Spain, 2019. “In-house protein

kinase inhibitors against Leishmania. The energy metabolism of the parasite as an off-target”

(Abstract book, P15). Authors: Martínez de Iturrate, P.; Sebastián-Pérez, V.;

Nácher-Vázquez, M.; Abengózar, M.A.; Martínez, A.; Campillo, N.E.; Rivas, L.; Gil, C.

Page 184: Protein-kinase inhibitors as potential leishmanicidal drugs
Page 185: Protein-kinase inhibitors as potential leishmanicidal drugs