Next Article in Journal
Failure of an Approximately Six Week Course of Tafenoquine to Completely Eradicate Babesia microti Infection in an Immunocompromised Patient
Next Article in Special Issue
Leishmaniasis Vaccines: Applications of RNA Technology and Targeted Clinical Trial Designs
Previous Article in Journal
Gut Microbiota and COVID-19: Potential Implications for Disease Severity
Previous Article in Special Issue
Morphogenesis Dynamics in Leishmania Differentiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Promastigote-to-Amastigote Conversion in Leishmania spp.—A Molecular View

Leishmania Genetics Group, Bernhard Nocht Institute for Tropical Medicine, D-20359 Hamburg, Germany
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Pathogens 2022, 11(9), 1052; https://doi.org/10.3390/pathogens11091052
Submission received: 31 May 2022 / Revised: 8 September 2022 / Accepted: 13 September 2022 / Published: 15 September 2022
(This article belongs to the Special Issue Leishmania & Leishmaniasis)

Abstract

:
A key factor in the successful infection of a mammalian host by Leishmania parasites is their conversion from extracellular motile promastigotes into intracellular amastigotes. We discuss the physical and chemical triggers that induce this conversion and the accompanying changes at the molecular level crucial for the survival of these intracellular parasites. Special emphasis is given to the reliance of these trypanosomatids on the post-transcriptional regulation of gene expression but also to the role played by protein kinases, chaperone proteins and proteolytic enzymes. Lastly, we offer a model to integrate the transduction of different stress signals for the induction of stage conversion.

1. Leishmania Biology and Life Cycle

1.1. Natural Life Cycle

All Leishmania parasites undergo a life cycle involving two different hosts, namely sand fly vectors (Diptera) and susceptible mammals, mostly rodents but also dogs and humans. To adapt and survive in their hosts, the leishmaniae exist in various flagellated forms inside the digestive tract of sand flies and as slowly proliferating or quiescent, non-motile amastigotes within macrophages and other mononuclear phagocytes.
Leishmania spp. are taken up by female flies into their digestive tracts as part of the blood meal. Within the peritrophic matrix, the parasites convert into flagellated promastigotes, which attach to the midgut epithelium via Leishmania surface molecules, e.g., lipophosphoglycan. These procyclic promastigotes proliferate logarithmically until they reach a stationary phase and undergo differentiation into small actively motile forms, the metacyclic promastigotes. The latter detach from the midgut epithelium and infect another mammal during the sand fly’s next bloodmeal [1,2]. Recent data also suggest that the Leishmania (Mundinia) subgenus may be transmitted by midges (genus Culicoides) [3]. Procyclic promastigotes of most Leishmania species can be generated in tissue culture media.
In mouse models, the injected parasites attract neutrophil granulocytes and tissue macrophages [4]. Inside these cells, leishmaniae convert into non-motile amastigotes, which persist in phagosomes, averting anti-microbial mechanisms by exporting effector molecules, i.e., regulatory proteins. They interfere with the signal transduction pathways of the host cells, modulating the host’s immune response [5].

1.2. Promastigote to Amastigote Differentiation

1.2.1. Temperature and pH as Triggers

Exposure of promastigotes to elevated temperatures and acidic milieu in vitro triggers their differentiation into amastigotes in axenic culture [6,7], facilitating the analysis of key molecular events during Leishmania stage conversion in this direction. The trigger temperatures must be adjusted according to the tissue tropism (viscera vs. skin) of the given Leishmania species and strains used, e.g., 37 °C for viscerotropic L. donovani and L. infantum and 32–34 °C for cutanotropic L. mexicana and L. (V.) braziliensis. This differentiation takes ~5 days to complete and is reversible [8]. Reverse conversion of axenic amastigotes back to promastigotes takes ~2 days to complete once the incubation temperature is lowered to ~25 °C and the pH returned to 7.0–7.4, mimicking conditions inside the sand fly gut.

1.2.2. Emerging Alternative/Additional Triggers

The majority of the biochemical and morphological changes seen in promastigote-to-amastigote differentiation are also observed when promastigotes are exposed to geldanamycin or radicicol, which occupy and inhibit the ATP-binding domain of the major chaperone protein, 90 kDa heat shock protein (HSP90 a.k.a. HSP83) [9,10].
There is also evidence for an involvement of iron metabolism and reactive oxygen species (ROS) in triggering the promastigote-to-amastigote conversion. Iron depletion in the growth media induces an increased Leishmania iron transporter 1 (LIT1) expression and stops proliferation, resulting in the appearance of an amastigote-like morphology and increased infectivity to susceptible mice [11]. This is accompanied by a higher abundance of certain amastigote-specific markers, such as amastin mRNAs. Exposure of Leishmania species to ROS was reported to produce a similar outcome [11,12].

1.2.3. Markers of Promastigote-to-Amastigote Conversion

Apart from the morphological changes, i.e., cell rounding and the reduction of the flagellum length, detectable by microscopy, there are molecular markers to verify stage conversion, e.g., amastigote-specific A2 proteins [13,14]. They exist as a family of proteins with variable copies of 10-amino acid repeats and are associated with Leishmania infectivity. Their primary role, however, appears to be stress protection [15,16].
Another amastigote-specific marker was identified in a proteomics study of L. donovani differentiation, i.e., 2,3-trans-enoyl CoA isomerase [17]. There are two isoforms of this enzyme discernible by SDS-PAGE and Western blot analysis. Isoform 1 and isoform 2 are expressed more abundantly in promastigotes and amastigotes, respectively [8]. Nuclease P4 may be amastigote-specific [18], as much as amastins [19], but its stage-specificity is equivocal due to a lack of common standard among different laboratories to verify Leishmania differentiation.

1.3. Limitations of In Vitro Differentiation Models

While axenically grown amastigotes are a very important tool for studying the dynamics of gene expression, the axenic conditions limit their utility for other fields of research, which require exposure to the host cell environments, i.e., antimicrobial defense mechanisms, residence in parasitophorous vacuoles and nutrient supplies from infected macrophages. This is reflected in the finding that proteins that are important for intracellular survival, such as the 100 kDa heat shock protein [20] or cyclophilin 40 [21], are dispensable for axenically grown amastigotes. The metabolic pathways are reprogrammed during Leishmania differentiation into axenic amastigotes, as reflective of the requirements of carbon sources from mammalian versus insect hosts and also the oxidative stress encountered in the former [17,22,23].
The use of axenic amastigotes for drug screening also raises the question of their appropriateness, considering that intracellular amastigotes protect themselves not only by the barrier of their own plasma membranes and detoxifying mechanisms but also by those of the host cells and parasitophorous vacuoles. Therefore, in vitro-infected cell lines or primary host cells are more appropriate for drug testing to provide more meaningful results [24,25].
While many Leishmania strains of cutanotropic and viscerotropic species readily convert into axenic amastigotes, L. major is a notable exception. A protocol was developed for the axenization of this species [26], but it has not been widely used to assess, for example, the effects of iron depletion and exposure to ROS [11].

2. Transduction of Differentiation Triggers

2.1. Post-Transcriptional Gene Regulation during Stage Conversion

Assessing gene expression in response to triggers of differentiation begins with the understanding of Leishmania’s uniquely structured genome, which is organized into polycistronic transcription units (PTU), each consisting of functionally unrelated, single-copied and/or tandemly repeated genes [27]. RNA polymerase II-dependent transcription proceeds through PTUs in the absence of canonical promoters and transcription factors and therefore lacks single-gene-specific regulation [28]. Instead, our current knowledge suggests that the PTUs are constitutively transcribed, producing polycistronic pre-mRNAs [29]. These pre-mRNAs are subsequently or, likely, co-transcriptionally processed by the linked events of trans-splicing, resulting in 5′ miniexon capping and 3′ polyadenylation to produce mature, monocistronic mRNAs [30]. During conversion from promastigotes to amastigotes, Leishmania must rapidly change the products of their gene expression, presumably regulated primarily by post-transcriptional mechanisms, for adaptation to the changing environment [31].
Leishmania gene expression is modulated, at least in part, by the stability of transcripts that are available for protein synthesis. Early studies suggest that Leishmania mRNA stability is controlled by cis-elements within the 5’- and 3′ untranslated regions (UTRs). The binding of RNA-associated proteins to these regulatory RNA motifs affects not only the stability of mRNAs but also their transport and intracellular targeting [32,33]. 3’UTR cis-elements have been associated with the stability and abundance of HSP mRNAs, which play an important role in adaptation to the increased temperature during amastigote differentiation [34]. Stage-specific up-regulation of the amastigote-specific surface protein amastin was initially thought to be regulated via cis-elements in the 3′UTRs of amastin mRNAs [35] but was subsequently found to occur at the translational stage [23,36].
In general, promastigote-to-amastigote differentiation is associated with a marked overall down-regulation of gene expression, especially for transcripts involved in translation and ribosome biogenesis, which is suggestive of a regulatory role [37]. This is reflected in the lower rate of translation seen in axenically differentiated amastigotes of L. infantum, resulting from phosphorylation of a translation factor called eukaryotic initiation factor 2. However, the translation of amastigote-specific proteins, such as the stress protein A2, is upregulated [38]. Leishmania differentiation from promastigote to amastigote is associated with a switch from cap-dependent to alternative translation, involving a non-conserved 4E-interacting protein (Leish4E-IP) [39].
In addition, the processing of policystronic to monomeric mRNAs has been suggested to regulate their abundance, hence controlling Leishmania stage-specific differentiation [40]. RNA processing entails the trans-splicing of the 39-bp spliced leader (SL) RNA onto the 5’ end and the addition of a poly(A) tail to the 3’ end of all matured mRNAs. However, this early hypothesis has not been confirmed, as the processing of individual RNAs does not have a significant impact on their steady state levels. In spite of notable differences in RNA processing between procyclic and metacyclic promastigotes [41], comparative RNA-Seq and ribosome profiling analyses revealed no significant correlation between changes in mRNA levels and protein synthesis rates [23].
Furthermore, RNA synthesis rates do not appear to have any influence on the adapta- tion of Leishmania to changing environmental conditions during differentiation. In early nuclear run on analyses [42], heat shock was found to have no effect on HSP mRNA synthesis. We are currently investigating RNA synthesis rates during stage conversion using precision nuclear run-on/sequencing analysis (PRO-seq) [43]. Initial data analyses suggest that RNA synthesis proceeds through intergenic sequences and is specific to the positive strand of each PTU. No stage-specific changes in RNA polymerase occupancy are observed in the PTUs (J.G., unpublished data).

2.2. Protein Turnover and Leishmania Proteases in Stage Differentiation

Leishmania conversion from elongated, flagellated promastigotes into ovoid amastigotes with rudimentary flagellum requires a retooled metabolism, resulting in multiple qualitative and quantitative proteomic changes, some of which are facilitated by proteolytic events. Leishmania proteases are involved in three key processes of amastigote conversion: (i) controlling the abundance of stage-specific proteins; (ii) inducing apoptosis to reduce parasite loads in the host cells; and (iii) modulating the host’s immune response.
The activities of Leishmania peptidases are regulated by the level of their expression, modification or stability and impact on the abundance of their substrate proteins. This is reflected in the stage-specific expression of many peptidases [17,23,44,45]. Accordingly, analysis of the serine proteases LbOBP and LbS13 [46] revealed that their expression is temperature-dependent, probably via 3’UTR secondary structure motifs and PTM [47].
Both proteasomal and lysosomal pathways are involved in Leishmania protein degradation [48]. During metacyclogenesis, cysteine protease B (CPB) levels increase in MVT-lysosomal compartments matching their proteolytic activities [48]. Acidification induces the secretion of LbCPB by L. braziliensis [49], and its specific activity peaks during the first 72 h of heat exposure in vitro [50]. The CPB is then secreted via the flagellar pocket.
The development of megasomes in amastigotes is linked to their infectivity [51]. The cysteine proteases therein are crucial for the parasites’ intracellular survival [48,52]. This is most likely due to the selective degradation of MHC II surface immune complexes in the infected macrophages and cytokines in the extracellular milieu, which shifts the host’s immune response toward the Th2 pathway [53].
A broad variety of stage-dependent functions is reported for the highly abundant and stage-specifically expressed metalloproteases of the MSP family (GP63, leishmanoly- sin). Their proposed functions include: (i) making nutrients available to Leishmania in the sandfly vector; (ii) releasing metacyclics from adhesion to the fly midgut epithelium; (iii) neutralizing the host’s first line antimicrobial defense activities, e.g., complement-mediated lysis, anti-microbial proteins and NK-cell attacks; (iv) facilitating Leishmania entry into macrophages by receptor-mediated phagocytosis and their motility in the extracellular matrix; and (v) maintaining the amastigotes’ survival in the parasitophorous vacuoles [54].
Leishmania differentiation is accompanied by autophagic events. Stressors that induce stage differentiation also trigger apoptosis via a variety of different molecular pathways [55]. In contrast to higher eukaryotes, Leishmania are deficient of caspases. Instead, some leishmaniae rely on metacaspases. This was demonstrated by inducing programmed cell death (PCD) in transgenic yeast by the over expression of LmjMCA [56]. Other Leishmania proteases of functional importance are a large family of calpain-like proteases (CALP) and cysteine protease C (CPC) [55]. The mechanisms employed by Leishmania to avoid apoptosis and instead enter differentiation deserve further detailed investigation, as this has the potential to provide novel targets for specific therapeutic intervention against leishmaniasis.

2.3. Protein Kinases in Stage Differentiation and Intracellular Survival

With Leishmania lacking transcriptional control, protein modifications constitute an important aspect of regulation [57]. A kinome-wide gene deletion study revealed that 162 L. mexicana protein kinase genes were dispensable, while 44 genes were refractory to deletion and thus considered essential for promastigote viability. In addition, 29 kinases were found to be crucial for differentiation from the metacyclic promastigote to the amastigote and for the successful infection of macrophages in vitro and BALB/c mice [58].
These kinases are highly conserved among different trypanosomatid species (Table 1) [58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84]. Moreover, some essential kinases were also found to play a crucial role in stage conversion and virulence, e.g., LmxMPK4 [85] and secreted CK1.2 [86]. The complete absence of tyrosine kinases, TKL receptors and RGC (receptor guanylyl cyclases) in trypanosomatids underscores the potential significance of Leishmania serine-/threonine kinases and of atypical kinase [58,62,87].
Leishmania kinases appear to mediate a wide variety of core functions in Leishmania differentiation, such as heat shock and osmotic stress response, nutrient sensing and metabolic regulation, translation control, cytokinesis, motility and morphogenesis (Table 1). Interestingly, the regulatory functions of Leishmania kinases often differ from those of their orthologues in higher eukaryotes. This complicates the efforts to predict interaction pathways. For instance, Leishmania appear to lack the activation cascade of LmxMPK4 by STE kinases at the conserved TxY Motif [58,75].
Due to their apparently crucial roles, Leishmania protein kinases also constitute promising drug targets notwithstanding their possible side effects against the host orthologues. Yet, a recent kinome comparison revealed a parasite-host sequence similarity of 23–69% (L. infantum vs. H. sapiens) and 21–69% (L. braziliensis vs. H. sapiens) [87], raising the hope of finding inhibitors with sufficient selectivity. Re-purposing available inhibitor libraries for testing against parasite kinases may be a promising path to finding lead compounds for treating Leishmania infections.

2.4. Epigenetic Effects

The structure of eukaryotic chromatin is primarily formed by nucleosomes, consisting of genomic DNA wrapped around a core histone octamer and separated by various lengths of linker DNA. This beads-on-a-string structure or 10-nm fibre is also referred to as euchromatin and represents accessible DNA. Further condensation leads to the formation of a 30-nm fibre—heterochromatin—which is considered less accessible for transcription factors and RNA polymerases. It has been postulated that euchromatin is required for inducing gene expression under changing environmental conditions, for example, during Leishmania stage conversion [88].
Indeed, ribosome profiling analyses showed that the inhibition of HSP90 leads promastigotes to become amastigote-like, resulting in the elevated synthesis of histones [23]. This was also observed by proteome analyses and occurs early on in the promastigote-to-amastigote conversion [17]. In addition, histone deacetylases were found to be regulated stage-dependently, suggesting the impact of epigenetic mechanisms on stage conversion [37]. An assay for transposase-accessible chromatin by sequencing (ATAC-seq) analysis showed that axenic amastigote differentiation is associated with the emergence of heterochromatin [89] in the divergent SSR regions upstream of telomere ends where transcription is initiated [90,91]. Whether the formation of heterochromatin in upstream SSRs has an impact on transcription rates in the downstream PTUs is currently under investigation (J.G., unpublished data).

2.5. HSPs in Stage Conversion

Since one of the key stimuli for promastigote-to-amastigote conversion is an elevation of the ambient temperature to that of the mammalian host, we suspected early on the involvement of the heat shock response in the development of amastigotes. The expression of various heat shock proteins is upregulated when subjecting promastigotes to a temperature up-shift, independent of acidic pH [92]. Therefore, the elevated synthesis of HSPs is an early response to this differentiation stimulus alone.
Moreover, the chaperones HSP100 and HSP23, along with the putative co-chaperone cyclophilin 40, were found to be essential for Leishmania survival in macrophages in vitro and/or in animal hosts [93,94,95]. HSP100 is functionally important to counter the host’s immune response in the parasite’s favor by playing a pivotal role in the assembly of immune-modulatory exosomes, as shown by their ineffectiveness when produced by HSP100 null mutants. While “wild type” exosomes trigger a cytokine expression in keeping with a Th2-response and a B cell-based immune reaction, the exosomes of HSP100 null mutants fail to suppress a Th1-based inflammatory response, known to restrict Leishmania survival in the mammalian host.
This is apparently due to the alteration in the composition of exosomes shed by HSP100−/− mutants lacking several HSPs and virulence factors [96].
The highly abundant chaperone HSP90 is involved in the transduction of differentiation signals. We first noted this when L. donovani promastigotes were treated with HSP90-specific inhibitors, geldanamycin or radicicol, resulting in a reduced growth and a morphological shift [9] from the long, slender, flagellated promastigotes to ovoid, non-motile amastigote-like cells, as shown in Figure 1.
Moreover, similar proteome changes were observed in axenic amastigotes and after radicicol- or 17-AAG-treatment [10]. Using ribosome profiling and BONCAT-iTRAQ mass spectrometry, respectively, the results showed elevated synthesis for a number of different proteins, including amastins, histones, fatty acid metabolic proteins, oxidative stress response proteins and several heat shock proteins, highly reminiscent of gene expression in axenic amastigotes generated by elevated temperature and acidity [23,97].
How HSP90 channels the signals to trigger Leishmania differentiation is unclear. It is known in yeast, plants and mammals that HSP90 and HSP70 act as negative regulators of the cellular heat shock response by quenching the activity of the HSF1 heat shock transcription factor [98]. There is little or no HSF activity detectable in homeostatic quiescent cells, but protein biosynthesis is not required for the rapid onset of HSF DNA-binding activity and heat shock gene expression [99]. Heat stress and protein folding stresses deplete the pool of free, active HSP70 and HSP90 chaperones, freeing HSF1 to form active trimers for binding its target sequences.
A similar depletion of HSP90 and HSP70 may occur during the exposure of Leishmania promastigotes to host tissue temperatures, a >∆10 °C heat shock compared with the sandfly gut. This may be enhanced by acidic milieu but also by iron depletion and ROS stress. Elevated temperature also boosts the shedding of exosomes which contain HSP90 and HSP70 as major payload proteins [100]. Such a depletion will cause HSP90-dependent regulatory proteins to be suppressed, even more so as HSP70 is also a part of the growth-promoting HSP90 multichaperone complex [101]. We speculate that HSP90 functional depletion transduces various stresses that trigger Leishmania stage conversion in vitro (Figure 2).
In higher eukaryotes, highly abundant HSP90 homologues are subject to phosphorylation with the concomitant modulation of its activity, but this requires the participation of protein kinases in an active or activatable state, such as MAP kinases [102,103,104]. Leishmania HSP90 is also the target of amastigote stage-specific phosphorylation, together with other housekeeping chaperones such as HSP70 and Sti1 [71]. HSP90 harbors several known or putative phosphorylation sites. Mutations of these sites produce a variety of different effects, ranging from the loss of cell viability to defective amastigote-specific infectivity (T223, S594, S595) to a minor reduction in proliferation [83].
So far, two protein kinases were identified that target HSP90: (i) MAP kinase 1, which was found to be essential for L. mexicana intracellular survival [60,81], and (ii) casein kinase 1, isoform 2 (CK1.2), which is secreted into the infected host cells and is required for intracellular survival [83,105]. The mutation of the CK1.2 phorphorylation site at S289 in HSP90 results in a slowdown of in vitro proliferation and minor morphological changes [83]. CK1.2 also interacts with HSP23 in vitro and can mediate stress resistance in the absence of HSP23 [106], while MAP kinase 1 also targets HSP70 [81].

3. Future Directions of Research

Recent years have seen a massive increase in experimental data about Leishmania stage conversion, mostly due to the application of various systems biology strategies aimed at the various levels of gene expression and their regulation. Still, there are pieces missing from the complex jigsaw puzzle on which we are working.
It is clear by now that mRNA abundance is relevant to, but not a reliable measure of, Leishmania gene expression. Since Leishmania, like all other trypanosomatids, regulate gene expression post-transcriptionally, it is more promising to direct our attention to the mechanisms of RNA utilization by ribosomes and translation factors. Still, a systems biology approach to correlate RNA stability with RNA abundance is expected to help explain the fluctuations of mRNA steady state levels observed in transcriptome analyses. Leishmania stage-specific gene expression is regulated primarily at the level of translational efficiency via the interactions of RNA-binding proteins with mRNAs for their stability and effective processing.
Proteome and translation analyses have shown that the expression patterns of individual ribosomal proteins undergo changes, suggestive of alterations in the composition and specificity of ribosomes during Leishmania stage conversion. Specific stress ribosomes have been reported to form under environmentally challenging conditions, leading to the preferential translation of specific mRNAs in bacteria. Post-translational protein modifications mediated chiefly by the activities of protein kinases may regulate the activity and specificity of translation initiation and elongation, signifying the need for investigation in this area and in the interactions between protein kinases and chaperones. It is essential to standardize the inducers of Leishmania stage conversion in vitro for examining the downstream pathways at the translatome and (phospho-) proteome levels to develop a unified model of signal transduction regulating this cyclic differentiation. The outcome of such investigation with the in vitro axenic differentiation model requires verification in vivo, that is now possible with the progress in single-cell sequencing, to study intracellular amastigotes and sand fly gut-derived promastigotes by NGS-based gene expression assays.

Author Contributions

Conceptualization, J.C., J.G. and M.H.; writing—original draft preparation, J.C., J.G. and M.H.; writing—review and editing, J.C., J.G. and M.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they are not aware of any conflicts of interest pertaining to this manuscript.

References

  1. Sacks, D.L.; Perkins, P.V. Development of infective stage Leishmania promastigotes within phlebotomine sand flies. Am. J. Trop. Med. Hyg. 1985, 34, 456–459. [Google Scholar] [CrossRef] [PubMed]
  2. Gossage, S.M.; Rogers, M.E.; Bates, P.A. Two separate growth phases during the development of Leishmania in sand flies: Implications for understanding the life cycle. Int. J. Parasitol. 2003, 33, 1027–1034. [Google Scholar] [CrossRef]
  3. Becvar, T.; Vojtkova, B.; Siriyasatien, P.; Votypka, J.; Modry, D.; Jahn, P.; Bates, P.; Carpenter, S.; Volf, P.; Sadlova, J. Experimental transmission of Leishmania (Mundinia) parasites by biting midges (Diptera: Ceratopogonidae). PLoS Pathog. 2021, 17, e1009654. [Google Scholar] [CrossRef] [PubMed]
  4. Laskay, T.; van Zandbergen, G.; Solbach, W. Neutrophil granulocytes—Trojan horses for Leishmania major and other intracellular microbes? Trends Microbiol. 2003, 11, 210–214. [Google Scholar] [CrossRef]
  5. Lambertz, U.; Silverman, J.M.; Nandan, D.; McMaster, W.R.; Clos, J.; Foster, L.J.; Reiner, N.E. Secreted virulence factors and immune evasion in visceral leishmaniasis. J. Leukoc. Biol. 2012, 91, 887–899. [Google Scholar] [CrossRef] [PubMed]
  6. Bates, P.A. Complete developmental cycle of Leishmania mexicanain axenic culture. Parasitology 1994, 108, 1–9. [Google Scholar] [CrossRef]
  7. Barak, E.; Amin-Spector, S.; Gerliak, E.; Goyard, S.; Holland, N.; Zilberstein, D. Differentiation of Leishmania donovani in host-free system: Analysis of signal perception and response. Mol. Biochem. Parasitol. 2005, 141, 99–108. [Google Scholar] [CrossRef]
  8. Zilberstein, D.; Nitzan Koren, R. Host-Free Systems for Differentiation of Axenic Leishmania. Methods Mol. Biol. 2019, 1971, 1–8. [Google Scholar] [CrossRef]
  9. Wiesgigl, M.; Clos, J. Heat Shock Protein 90 Homeostasis Controls Stage Differentiation in Leishmania donovani. Mol. Biol. Cell 2001, 12, 3307–3316. [Google Scholar] [CrossRef]
  10. Bente, M.; Harder, S.; Wiesgigl, M.; Heukeshoven, J.; Gelhaus, C.; Krause, E.; Clos, J.; Bruchhaus, I. Developmentally induced changes of the proteome in the protozoan parasite Leishmania donovani. Proteomics 2003, 3, 1811–1829. [Google Scholar] [CrossRef]
  11. Mittra, B.; Cortez, M.; Haydock, A.; Ramasamy, G.; Myler, P.J.; Andrews, N.W. Iron uptake controls the generation of Leishmania infective forms through regulation of ROS levels. J. Exp. Med. 2013, 210, 401–416. [Google Scholar] [CrossRef] [PubMed]
  12. Khan, Y.A.; Andrews, N.W.; Mittra, B. ROS regulate differentiation of visceralizing Leishmania species into the virulent amastigote form. Parasitol. Open 2018, 4, e19. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, W.W.; Charest, H.; Ghedin, E.; Matlashewski, G. Identification and overexpression of the A2 amastigote-specific protein in Leishmania donovani. Mol. Biochem. Parasitol. 1996, 78, 79–90. [Google Scholar] [CrossRef]
  14. Zhang, W.W.; Matlashewski, G. Loss of virulence in Leishmania donovani deficient in an amastigote-specific protein, A2. Proc. Natl. Acad. Sci. USA 1997, 94, 8807–8811. [Google Scholar] [CrossRef] [PubMed]
  15. McCall, L.I.; Matlashewski, G. Localization and induction of the A2 virulence factor in Leishmania: Evidence that A2 is a stress response protein. Mol. Microbiol. 2010, 77, 518–530. [Google Scholar] [CrossRef]
  16. McCall, L.I.; Matlashewski, G. Involvement of the Leishmania donovani virulence factor A2 in protection against heat and oxidative stress. Exp. Parasitol. 2012, 132, 109–115. [Google Scholar] [CrossRef]
  17. 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. [Google Scholar] [CrossRef]
  18. Kar, S.; Soong, L.; Colmenares, M.; Goldsmith-Pestana, K.; McMahon-Pratt, D. The immunologically protective P-4 antigen of Leishmania amastigotes. A developmentally regulated single strand-specific nuclease associated with the endoplasmic reticulum. J. Biol. Chem. 2000, 275, 37789–37797. [Google Scholar] [CrossRef]
  19. Rochette, A.; McNicoll, F.; Girard, J.; Breton, M.; Leblanc, E.; Bergeron, M.G.; Papadopoulou, B. Characterization and developmental gene regulation of a large gene family encoding amastin surface proteins in Leishmania spp. Mol. Biochem. Parasitol. 2005, 140, 205–220. [Google Scholar] [CrossRef]
  20. Krobitsch, S.; Clos, J. A novel role for 100 kD heat shock proteins in the parasite Leishmania donovani. Cell Stress Chaperones 1999, 4, 191–198. [Google Scholar] [CrossRef]
  21. Yau, W.L.; Pescher, P.; MacDonald, A.; Hem, S.; Zander, D.; Retzlaff, S.; Blisnick, T.; Rotureau, B.; Rosenqvist, H.; Wiese, M.; et al. The Leishmania donovani chaperone cyclophilin 40 is essential for intracellular infection independent of its stage-specific phosphorylation status. Mol. Microbiol. 2014, 93, 80–97. [Google Scholar] [CrossRef] [PubMed]
  22. 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. [Google Scholar] [CrossRef] [PubMed]
  23. Bifeld, E.; Lorenzen, S.; Bartsch, K.; Vasquez, J.J.; Siegel, T.N.; Clos, J. Ribosome Profiling Reveals HSP90 Inhibitor Effects on Stage-Specific Protein Synthesis in Leishmania donovani. mSystems 2018, 3. [Google Scholar] [CrossRef] [PubMed]
  24. Moraes, C.B.; Alcantara, L.M. Quantification of Parasite Loads by Automated Microscopic Image Analysis. Methods Mol. Biol. 2019, 1971, 279–288. [Google Scholar] [CrossRef]
  25. Santarem, N.; Tavares, J.; Cordeiro-da-Silva, A. In Vitro Infections of Macrophage-Like Cell Lines with Leishmania infantum for Drug Screening. Methods Mol. Biol. 2019, 1971, 265–277. [Google Scholar] [CrossRef] [PubMed]
  26. Wenzel, U.A.; Bank, E.; Florian, C.; Forster, S.; Zimara, N.; Steinacker, J.; Klinger, M.; Reiling, N.; Ritter, U.; van Zandbergen, G. Leishmania major parasite stage-dependent host cell invasion and immune evasion. FASEB J. 2012, 26, 29–39. [Google Scholar] [CrossRef]
  27. Ivens, A.C.; Peacock, C.S.; Worthey, E.A.; Murphy, L.; Aggarwal, G.; Berriman, M.; Sisk, E.; Rajandream, M.A.; Adlem, E.; Aert, R.; et al. The genome of the kinetoplastid parasite, Leishmania major. Science 2005, 309, 436–442. [Google Scholar] [CrossRef]
  28. Clayton, C.E. Life without transcriptional control? From fly to man and back again. EMBO J. 2002, 21, 1881–1888. [Google Scholar] [CrossRef]
  29. Leifso, K.; Cohen-Freue, G.; Dogra, N.; Murray, A.; McMaster, W.R. Genomic and proteomic expression analysis of Leishmania promastigote and amastigote life stages: The Leishmania genome is constitutively expressed. Mol. Biochem. Parasitol. 2007, 152, 35–46. [Google Scholar] [CrossRef]
  30. Myler, P.J.; Beverley, S.M.; Cruz, A.K.; Dobson, D.E.; Ivens, A.C.; McDonagh, P.D.; Madhubala, R.; Martinez-Calvillo, S.; Ruiz, J.C.; Saxena, A.; et al. The Leishmania genome project: New insights into gene organization and function. Med. Microbiol. Immunol. 2001, 190, 9–12. [Google Scholar] [CrossRef]
  31. Hunter, K.W.; Cook, C.L.; Hayunga, E.G. Leishmanial differentiation in vitro: Induction of heat shock proteins. Biochem. Biophys. Res. Commun. 1984, 125, 755–760. [Google Scholar] [CrossRef]
  32. Aly, R.; Argaman, M.; Halman, S.; Shapira, M. A regulatory role for the 5’ and 3’ untranslated regions in differential expression of hsp83 in Leishmania. Nucleic Acids Res. 1994, 22, 2922–2929. [Google Scholar] [CrossRef] [PubMed]
  33. Nandan, D.; Thomas, S.A.; Nguyen, A.; Moon, K.M.; Foster, L.J.; Reiner, N.E. Comprehensive Identification of mRNA-Binding Proteins of Leishmania donovani by Interactome Capture. PLoS ONE 2017, 12, e0170068. [Google Scholar] [CrossRef] [PubMed]
  34. Larreta, R.; Soto, M.; Quijada, L.; Folgueira, C.; Abanades, D.R.; Alonso, C.; Requena, J.M. The expression of HSP83 genes in Leishmania infantum is affected by temperature and by stage-differentiation and is regulated at the levels of mRNA stability and translation. BMC Mol. Biol. 2004, 5, 3. [Google Scholar] [CrossRef]
  35. Wu, Y.; El Fakhry, Y.; Sereno, D.; Tamar, S.; Papadopoulou, B. A new developmentally regulated gene family in Leishmania amastigotes encoding a homolog of amastin surface proteins. Mol. Biochem. Parasitol. 2000, 110, 345–357. [Google Scholar] [CrossRef]
  36. McNicoll, F.; Muller, M.; Cloutier, S.; Boilard, N.; Rochette, A.; Dube, M.; Papadopoulou, B. Distinct 3’-untranslated region elements regulate stage-specific mRNA accumulation and translation in Leishmania. J. Biol. Chem. 2005, 280, 35238–35246. [Google Scholar] [CrossRef]
  37. Alcolea, P.J.; Alonso, A.; Gomez, M.J.; Moreno, I.; Dominguez, M.; Parro, V.; Larraga, V. Transcriptomics throughout the life cycle of Leishmania infantum: High down-regulation rate in the amastigote stage. Int. J. Parasitol. 2010, 40, 1497–1516. [Google Scholar] [CrossRef]
  38. Cloutier, S.; Laverdiere, M.; Chou, M.N.; Boilard, N.; Chow, C.; Papadopoulou, B. Translational control through eIF2alpha phosphorylation during the Leishmania differentiation process. PLoS ONE 2012, 7, e35085. [Google Scholar] [CrossRef]
  39. Zinoviev, A.; Leger, M.; Wagner, G.; Shapira, M. A novel 4E-interacting protein in Leishmania is involved in stage-specific translation pathways. Nucleic Acids Res. 2011, 39, 8404–8415. [Google Scholar] [CrossRef]
  40. Rastrojo, A.; Carrasco-Ramiro, F.; Martin, D.; Crespillo, A.; Reguera, R.M.; Aguado, B.; Requena, J.M. The transcriptome of Leishmania major in the axenic promastigote stage: Transcript annotation and relative expression levels by RNA-seq. BMC Genom. 2013, 14, 223. [Google Scholar] [CrossRef] [Green Version]
  41. Dillon, L.A.; Okrah, K.; Hughitt, V.K.; Suresh, R.; Li, Y.; Fernandes, M.C.; Belew, A.T.; Corrada Bravo, H.; Mosser, D.M.; El-Sayed, N.M. Transcriptomic profiling of gene expression and RNA processing during Leishmania major differentiation. Nucleic Acids Res. 2015, 43, 6799–6813. [Google Scholar] [CrossRef] [PubMed]
  42. Brandau, S.; Dresel, A.; Clos, J. High constitutive levels of heat-shock proteins in human-pathogenic parasites of the genus Leishmania. Biochem. J. 1995, 310 Pt 2, 225–232. [Google Scholar] [CrossRef]
  43. Mahat, D.B.; Kwak, H.; Booth, G.T.; Jonkers, I.H.; Danko, C.G.; Patel, R.K.; Waters, C.T.; Munson, K.; Core, L.J.; Lis, J.T. Base-pair-resolution genome-wide mapping of active RNA polymerases using precision nuclear run-on (PRO-seq). Nat. Protoc. 2016, 11, 1455–1476. [Google Scholar] [CrossRef] [PubMed]
  44. Mottram, J.C.; Frame, M.J.; Brooks, D.R.; Tetley, L.; Hutchison, J.E.; Souza, A.E.; Coombs, G.H. The multiple cpb cysteine proteinase genes of Leishmania mexicana encode isoenzymes that differ in their stage regulation and substrate preferences. J. Biol. Chem. 1997, 272, 14285–14293. [Google Scholar] [CrossRef] [PubMed]
  45. Besteiro, S.; Williams, R.A.; Coombs, G.H.; Mottram, J.C. Protein turnover and differentiation in Leishmania. Int. J. Parasitol. 2007, 37, 1063–1075. [Google Scholar] [CrossRef]
  46. Zabala-Peñafiel, A.; Cysne-Finkelstein, L.; Conceição-Silva, F.; Fagundes, A.; Miranda, L.d.F.C.; Souza-Silva, F.; Brandt, A.A.M.L.; Dias-Lopes, G.; Alves, C.R. Novel Insights Into Leishmania (Viannia) braziliensis In Vitro Fitness Guided by Temperature Changes Along With Its Subtilisins and Oligopeptidase B. Front. Cell. Infect. Microbiol. 2022, 12, 411. [Google Scholar] [CrossRef]
  47. Ramu, D.; Singh, S. Potential molecular targets of Leishmania pathways in developing novel antileishmanials. Future Microbiol. 2022, 17, 41–57. [Google Scholar] [CrossRef]
  48. Waller, R.F.; McConville, M.J. Developmental changes in lysosome morphology and function Leishmania parasites. Int. J. Parasitol. 2002, 32, 1435–1445. [Google Scholar] [CrossRef]
  49. Ke, G.; Mauel, J.; Rivier, D. Leishmania mexicana: Extracellular proton concentration is a key regulator of cysteine proteinase CPb expression. Exp. Parasitol. 1998, 90, 58–64. [Google Scholar] [CrossRef]
  50. Gomes, C.B.; Silva, F.S.; Charret, K.D.; Pereira, B.A.; Finkelstein, L.C.; Santos-de-Souza, R.; de Castro Cortes, L.M.; Pereira, M.C.; Rodrigues de Oliveira, F.O., Jr.; Alves, C.R. Increasing in cysteine proteinase B expression and enzymatic activity during in vitro differentiation of Leishmania (Viannia) braziliensis: First evidence of modulation during morphological transition. Biochimie 2017, 133, 28–36. [Google Scholar] [CrossRef]
  51. Pral, E.M.; Moitinho, M.d.L.R.; Balanco, J.M.F.; Teixeira, V.R.; Milder, R.V.; Alfieri, S.C. Growth phase and medium pH modulate the expression of proteinase activities and the development of megasomes in axenically cultivated Leishmania (Leishmania) amazonensis amastigote–like organisms. J. Parasitol. 2003, 89, 35–43. [Google Scholar] [CrossRef]
  52. Ueda-Nakamura, T.; da Conceição Rocha Sampaio, M.; Cunha-e-Silva, N.L.; Traub-Cseko, Y.M.; de Souza, W. Expression and processing of megasome cysteine proteinases during Leishmania amazonensis differentiation. Parasitol. Res. 2002, 88, 332–337. [Google Scholar] [CrossRef] [PubMed]
  53. Courret, N.; Frehel, C.; Prina, E.; Lang, T.; Antoine, J.C. Kinetics of the intracellular differentiation of Leishmania amazonensis and internalization of host MHC molecules by the intermediate parasite stages. Parasitology 2001, 122, 263–279. [Google Scholar] [CrossRef] [PubMed]
  54. Yao, C. Major surface protease of trypanosomatids: One size fits all? Infect. Immun. 2010, 78, 22–31. [Google Scholar] [CrossRef]
  55. Basmaciyan, L.; Casanova, M. Cell death in Leishmania. Parasite 2019, 26, 71. [Google Scholar] [CrossRef]
  56. Gonzalez, I.J.; Desponds, C.; Schaff, C.; Mottram, J.C.; Fasel, N. Leishmania major metacaspase can replace yeast metacaspase in programmed cell death and has arginine-specific cysteine peptidase activity. Int. J. Parasitol. 2007, 37, 161–172. [Google Scholar] [CrossRef]
  57. Spath, G.F.; Drini, S.; Rachidi, N. A touch of Zen: Post-translational regulation of the Leishmania stress response. Cell Microbiol. 2015, 17, 632–638. [Google Scholar] [CrossRef]
  58. Baker, N.; Catta-Preta, C.M.C.; Neish, R.; Sadlova, J.; Powell, B.; Alves-Ferreira, E.V.C.; Geoghegan, V.; Carnielli, J.B.T.; Newling, K.; Hughes, C.; et al. Systematic functional analysis of Leishmania protein kinases identifies regulators of differentiation or survival. Nat. Commun. 2021, 12, 1244. [Google Scholar] [CrossRef]
  59. Fischer Weinberger, R.; Bachmaier, S.; Dandugudumula, R.; Phan, I.Q.; Almoznino, M.; Githure, G.B.; Polatoglou, E.; Tsigankov, P.; Nitzan Koren, R.; Myler, P.J.; et al. A divergent protein kinase A in the human pathogen Leishmania is associated with developmental morphogenesis. bioRxiv 2021. under review. [Google Scholar] [CrossRef]
  60. Wiese, M. A mitogen-activated protein (MAP) kinase homologue of Leishmania mexicana is essential for parasite survival in the infected host. EMBO J. 1998, 17, 2619–2628. [Google Scholar] [CrossRef] [Green Version]
  61. Agron, P.G.; Reed, S.L.; Engel, J.N. An essential, putative MEK kinase of Leishmania major. Mol. Biochem. Parasitol. 2005, 142, 121–125. [Google Scholar] [CrossRef] [PubMed]
  62. 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 Genom. 2005, 6, 127. [Google Scholar] [CrossRef] [PubMed]
  63. De Lima, A.R.; Medina, R.; Uzcanga, G.L.; Noris Suarez, K.; Contreras, V.T.; Navarro, M.C.; Arteaga, R.; Bubis, J. Tight binding between a pool of the heterodimeric alpha/beta tubulin and a protein kinase CK2 in Trypanosoma cruzi epimastigotes. Parasitology 2006, 132, 511–523. [Google Scholar] [CrossRef] [PubMed]
  64. Hammarton, T.C. Cell cycle regulation in Trypanosoma brucei. Mol. Biochem. Parasitol. 2007, 153, 1–8. [Google Scholar] [CrossRef] [PubMed]
  65. Moraes, M.C.; Jesus, T.C.; Hashimoto, N.N.; Dey, M.; Schwartz, K.J.; Alves, V.S.; Avila, C.C.; Bangs, J.D.; Dever, T.E.; Schenkman, S.; et al. Novel membrane-bound eIF2alpha kinase in the flagellar pocket of Trypanosoma brucei. Eukaryot. Cell 2007, 6, 1979–1991. [Google Scholar] [CrossRef]
  66. Morales, M.A.; Renaud, O.; Faigle, W.; Shorte, S.L.; Spath, G.F. Over-expression of Leishmania major MAP kinases reveals stage-specific induction of phosphotransferase activity. Int. J. Parasitol. 2007, 37, 1187–1199. [Google Scholar] [CrossRef]
  67. Dutra, P.M.; Vieira, D.P.; Meyer-Fernandes, J.R.; Silva-Neto, M.A.; Lopes, A.H. Stimulation of Leishmania tropica protein kinase CK2 activities by platelet-activating factor (PAF). Acta Trop. 2009, 111, 247–254. [Google Scholar] [CrossRef]
  68. Rotureau, B.; Morales, M.A.; Bastin, P.; Späth, G.F. The flagellum-mitogen-activated protein kinase connection in Trypanosomatids: A key sensory role in parasite signalling and development? Cell Microbiol. 2009, 11, 710–718. [Google Scholar] [CrossRef]
  69. Madeira da Silva, L.; Beverley, S.M. Expansion of the target of rapamycin (TOR) kinase family and function in Leishmania shows that TOR3 is required for acidocalcisome biogenesis and animal infectivity. Proc. Natl. Acad. Sci. USA 2010, 107, 11965–11970. [Google Scholar] [CrossRef]
  70. Morales, M.A.; Pescher, P.; Spath, G.F. Leishmania major MPK7 protein kinase activity inhibits intracellular growth of the pathogenic amastigote stage. Eukaryot. Cell 2010, 9, 22–30. [Google Scholar] [CrossRef] [Green Version]
  71. Morales, M.A.; Watanabe, R.; Dacher, M.; Chafey, P.; Osorio y Fortea, J.; Scott, D.A.; Beverley, S.M.; Ommen, G.; Clos, J.; Hem, S.; et al. Phosphoproteome dynamics reveal heat-shock protein complexes specific to the Leishmania donovani infectious stage. Proc. Natl. Acad. Sci. USA 2010, 107, 8381–8386. [Google Scholar] [CrossRef] [PubMed]
  72. Horjales, S.; Schmidt-Arras, D.; Limardo, R.R.; Leclercq, O.; Obal, G.; Prina, E.; Turjanski, A.G.; Spath, G.F.; Buschiazzo, A. The crystal structure of the MAP kinase LmaMPK10 from Leishmania major reveals parasite-specific features and regulatory mechanisms. Structure 2012, 20, 1649–1660. [Google Scholar] [CrossRef] [PubMed]
  73. Mandal, G.; Sharma, M.; Kruse, M.; Sander-Juelch, C.; Munro, L.A.; Wang, Y.; Vilg, J.V.; Tamas, M.J.; Bhattacharjee, H.; Wiese, M.; et al. Modulation of Leishmania major aquaglyceroporin activity by a mitogen-activated protein kinase. Mol. Microbiol. 2012, 85, 1204–1218. [Google Scholar] [CrossRef]
  74. Badjatia, N.; Ambrósio, D.L.; Lee, J.H.; Günzl, A. Trypanosome cdc2-related kinase 9 controls spliced leader RNA cap4 methylation and phosphorylation of RNA polymerase II subunit RPB1. Mol. Cell. Biol. 2013, 33, 1965–1975. [Google Scholar] [CrossRef]
  75. Cayla, M.; Rachidi, N.; Leclercq, O.; Schmidt-Arras, D.; Rosenqvist, H.; Wiese, M.; Spath, G.F. Transgenic analysis of the Leishmania MAP kinase MPK10 reveals an auto-inhibitory mechanism crucial for stage-regulated activity and parasite viability. PLoS Pathog. 2014, 10, e1004347. [Google Scholar] [CrossRef]
  76. Garg, M.; Goyal, N. MAPK1 of Leishmania donovani modulates antimony susceptibility by downregulating P-glycoprotein efflux pumps. Antimicrob. Agents Chemother. 2015, 59, 3853–3863. [Google Scholar] [CrossRef]
  77. Inoue, M.; Okamoto, K.; Uemura, H.; Yasuda, K.; Motohara, Y.; Morita, K.; Hiromura, M.; Reddy, E.P.; Fukuma, T.; Horikoshi, N. Identification and characterization of a cell division-regulating kinase AKB1 (associated kinase of Trypanosoma brucei 14-3-3) through proteomics study of the Tb14-3-3 binding proteins. J. Biochem. 2015, 158, 49–60. [Google Scholar] [CrossRef]
  78. 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. [Google Scholar] [CrossRef]
  79. Goldman-Pinkovich, A.; Balno, C.; Strasser, R.; Zeituni-Molad, M.; Bendelak, K.; Rentsch, D.; Ephros, M.; Wiese, M.; Jardim, A.; Myler, P.J.; et al. An Arginine Deprivation Response Pathway Is Induced in Leishmania during Macrophage Invasion. PLoS Pathog. 2016, 12, e1005494. [Google Scholar] [CrossRef]
  80. Fernandez-Cortes, F.; Serafim, T.D.; Wilkes, J.M.; Jones, N.G.; Ritchie, R.; McCulloch, R.; Mottram, J.C. RNAi screening identifies Trypanosoma brucei stress response protein kinases required for survival in the mouse. Sci. Rep. 2017, 7, 6156. [Google Scholar] [CrossRef] [Green Version]
  81. 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. [Google Scholar] [CrossRef] [PubMed]
  82. Varga, V.; Moreira-Leite, F.; Portman, N.; Gull, K. Protein diversity in discrete structures at the distal tip of the trypanosome flagellum. Proc. Natl. Acad. Sci. USA 2017, 114, E6546–E6555. [Google Scholar] [CrossRef] [PubMed]
  83. Hombach-Barrigah, A.; Bartsch, K.; Smirlis, D.; Rosenqvist, H.; MacDonald, A.; Dingli, F.; Loew, D.; Spath, G.F.; Rachidi, N.; Wiese, M.; et al. Leishmania donovani 90 kD Heat Shock Protein—Impact of Phosphosites on Parasite Fitness, Infectivity and Casein Kinase Affinity. Sci. Rep. 2019, 9, 5074. [Google Scholar] [CrossRef]
  84. Kelly, F.D.; Yates, P.A.; Landfear, S.M. Nutrient sensing in Leishmania: Flagellum and cytosol. Mol. Microbiol. 2021, 115, 849–859. [Google Scholar] [CrossRef] [PubMed]
  85. 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. [Google Scholar] [CrossRef]
  86. Liu, J.; Carvalho, L.P.; Bhattacharya, S.; Carbone, C.J.; Kumar, K.G.; Leu, N.A.; Yau, P.M.; Donald, R.G.; Weiss, M.J.; Baker, D.P.; et al. Mammalian casein kinase 1alpha and its leishmanial ortholog regulate stability of IFNAR1 and type I interferon signaling. Mol. Cell. Biol. 2009, 29, 6401–6412. [Google Scholar] [CrossRef] [PubMed]
  87. 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. [Google Scholar] [CrossRef] [PubMed]
  88. Belli, S.I. Chromatin remodelling during the life cycle of trypanosomatids. Int. J. Parasitol. 2000, 30, 679–687. [Google Scholar] [CrossRef]
  89. Grunebast, J.; Lorenzen, S.; Zummack, J.; Clos, J. Life Cycle Stage-Specific Accessibility of Leishmania donovani Chromatin at Transcription Start Regions. mSystems 2021, 6, e0062821. [Google Scholar] [CrossRef]
  90. Martinez-Calvillo, S.; Yan, S.; Nguyen, D.; Fox, M.; Stuart, K.; Myler, P.J. Transcription of Leishmania major Friedlin chromosome 1 initiates in both directions within a single region. Mol. Cell 2003, 11, 1291–1299. [Google Scholar] [CrossRef]
  91. Martinez-Calvillo, S.; Nguyen, D.; Stuart, K.; Myler, P.J. Transcription initiation and termination on Leishmania major chromosome 3. Eukaryot. Cell 2004, 3, 506–517. [Google Scholar] [CrossRef] [PubMed]
  92. Kröber-Boncardo, C.; Grünebast, J.; Clos, J. Heat Shock Proteins in Leishmania Parasites. In Heat Shock Proteins; Asea, A., Ed.; Springer: Dordrecht, The Netherlands, 2020. [Google Scholar] [CrossRef]
  93. Hubel, A.; Krobitsch, S.; Horauf, A.; Clos, J. Leishmania major Hsp100 is required chiefly in the mammalian stage of the parasite. Mol. Cell. Biol. 1997, 17, 5987–5995. [Google Scholar] [CrossRef] [PubMed]
  94. Hombach, A.; Ommen, G.; MacDonald, A.; Clos, J. A small heat shock protein is essential for thermotolerance and intracellular survival of Leishmania donovani. J. Cell Sci. 2014, 127, 4762–4773. [Google Scholar] [CrossRef]
  95. Yau, W.L.; Lambertz, U.; Colineau, L.; Pescher, P.; MacDonald, A.; Zander, D.; Retzlaff, S.; Eick, J.; Reiner, N.E.; Clos, J.; et al. Phenotypic Characterization of a Leishmania donovani Cyclophilin 40 Null Mutant. J. Eukaryot. Microbiol. 2016, 63, 823–833. [Google Scholar] [CrossRef]
  96. Silverman, J.M.; Clos, J.; Horakova, E.; Wang, A.Y.; Wiesgigl, M.; Kelly, I.; Lynn, M.A.; McMaster, W.R.; Foster, L.J.; Levings, M.K.; et al. Leishmania exosomes modulate innate and adaptive immune responses through effects on monocytes and dendritic cells. J. Immunol. 2010, 185, 5011–5022. [Google Scholar] [CrossRef]
  97. Kalesh, K.; Sundriyal, S.; Perera, H.; Cobb, S.L.; Denny, P.W. Quantitative Proteomics Reveals that Hsp90 Inhibition Dynamically Regulates Global Protein Synthesis in Leishmania mexicana. mSystems 2021, 6, e00089-21. [Google Scholar] [CrossRef]
  98. Masser, A.E.; Ciccarelli, M.; Andreasson, C. Hsf1 on a leash—Controlling the heat shock response by chaperone titration. Exp. Cell Res. 2020, 396, 112246. [Google Scholar] [CrossRef]
  99. Zimarino, V.; Wu, C. Induction of sequence specific binding of Drosophila heat shock activator protein without protein synthesis. Nature 1987, 327, 727–730. [Google Scholar] [CrossRef]
  100. Silverman, J.M.; Clos, J.; de’Oliveira, C.C.; Shirvani, O.; Fang, Y.; Wang, C.; Foster, L.J.; Reiner, N.E. An exosome-based secretion pathway is responsible for protein export from Leishmania and communication with macrophages. J. Cell Sci. 2010, 123, 842–852. [Google Scholar] [CrossRef] [Green Version]
  101. Ommen, G.; Chrobak, M.; Clos, J. The co-chaperone SGT of Leishmania donovaniis essential for the parasite’s viability. Cell Stress Chaperones 2010, 39, 541–546. [Google Scholar] [CrossRef]
  102. Pratt, W.B. The role of the hsp90-based chaperone system in signal transduction by nuclear receptors and receptors signaling via MAP kinase. Annu. Rev. Pharmacol. Toxicol. 1997, 37, 297–326. [Google Scholar] [CrossRef] [PubMed]
  103. Zhang, H.; Burrows, F. Targeting multiple signal transduction pathways through inhibition of Hsp90. J. Mol. Med. 2004, 82, 488–499. [Google Scholar] [CrossRef] [PubMed]
  104. Verma, S.; Goyal, S.; Jamal, S.; Singh, A.; Grover, A. Hsp90: Friends, clients and natural foes. Biochimie 2016, 127, 227–240. [Google Scholar] [CrossRef] [PubMed]
  105. 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 kinase 1 as a drug target and reveals important functions in parasite viability and intracellular infection. Antimicrob. Agents Chemother. 2014, 58, 1501–1515. [Google Scholar] [CrossRef]
  106. Kröber-Boncardo, C.; Lorenzen, S.; Brinker, C.; Clos, J. Casein kinase 1.2 over expression restores stress resistance to Leishmania donovani HSP23 null mutants. Sci. Rep. 2020, 10, 15969. [Google Scholar] [CrossRef]
Figure 1. Axenic stage conversion of L. donovani from promastigotes kept at 25 °C/pH 7.4. (A) to amastigotes at 37 °C/pH 5.5 (B) and chemically (radicicol)-induced amastigote-like forms (C). Samples were fixed and stained with DAPI (nuclei, blue) and anti-α-tubulin mAB (microtubuli, red). Fluorescence images were captured at a 100× magnification using an EVOS Autofluor microscope; overlays were implemented in Adobe Photoshop CS3. The bars represent 10 µm.
Figure 1. Axenic stage conversion of L. donovani from promastigotes kept at 25 °C/pH 7.4. (A) to amastigotes at 37 °C/pH 5.5 (B) and chemically (radicicol)-induced amastigote-like forms (C). Samples were fixed and stained with DAPI (nuclei, blue) and anti-α-tubulin mAB (microtubuli, red). Fluorescence images were captured at a 100× magnification using an EVOS Autofluor microscope; overlays were implemented in Adobe Photoshop CS3. The bars represent 10 µm.
Pathogens 11 01052 g001
Figure 2. Schematic model of stage conversion through HSP90 quenching. Axenically grown promastigotes (A) as elongated, flagellated, highly proliferative cells are exposed to cell stress (increased temperature, acidified medium, iron depletion or ROS). This causes the sequestration of HSP90 and other foldosome components to denatured protein domains (B). The temperature-dependent shedding of HSP-containing exosomes and the phosphorylation of HSP90 by CK1.2 and MAPK1 adds to the reduction of free HSP90, which triggers conversion to the ovoid, non-motile, growth-impaired amastigote (C). Alternatively, the direct inhibition of HSP90 with radicicol or geldanamycin also depletes active HSP90 and causes a similar conversion to amastigote-like stages (D).
Figure 2. Schematic model of stage conversion through HSP90 quenching. Axenically grown promastigotes (A) as elongated, flagellated, highly proliferative cells are exposed to cell stress (increased temperature, acidified medium, iron depletion or ROS). This causes the sequestration of HSP90 and other foldosome components to denatured protein domains (B). The temperature-dependent shedding of HSP-containing exosomes and the phosphorylation of HSP90 by CK1.2 and MAPK1 adds to the reduction of free HSP90, which triggers conversion to the ovoid, non-motile, growth-impaired amastigote (C). Alternatively, the direct inhibition of HSP90 with radicicol or geldanamycin also depletes active HSP90 and causes a similar conversion to amastigote-like stages (D).
Pathogens 11 01052 g002
Table 1. Leishmania protein kinases and their known roles/functions in stage differentiation.
Table 1. Leishmania protein kinases and their known roles/functions in stage differentiation.
FamilyKinaseReported FunctionReference
CMGC/CDKLmx/T.br/Tc/LmjCRK7No functional homologue to human CDK7Baker et al.,2021 [53]
Badjatia et al., 2013 [70]
Parsons et al., 2005 [58]
LmxCRK8Untypical regulation indicated in TrCRK8Baker et al., 2021 [53]
Hammarton 2007 [60]
CMGC/CDKL and CMGC/MAPKLmx/LdMPK1Intracellular survival, phosphorylating LdHSP70 and LdHSP90; antimony resistanceBaker et al., 2021 [53]
Wiese, 1998 [56]
Hombach-Barrigah et al., 2019 [79] Kaur et al., 2017 [77]
Morales et al., 2010b [67]
Garg and Goyal, 2015 [72]
LmxMPK2Infection; essential nutrient regulation and osmotic stress via Arginine depletion response (ADR) and AQP1-regulation; antimony resistance
flagella-mediated environment sensing
Goldman-Pinkovich et al., 2016 [75]
Kelly et al., 2021 [80]
Mandal et al., 2012 [69]
Rotureau et al. 2009 [64]
LmxMPK15InfectionBaker et al., 2021 [53]
Lmx/LmjMPK10Stage-specific auto-regulation and phosphorylation crucial for infection; crystal structure available;
not crucial in L. major
Morales et al., 2007 [62]
Horjales et al., 2012 [68]
Cayla et al., 2014 [71]
LmjMPK7InfectionMorales et al., 2010 [67]
STELmx/Lmj/T.brMRK1Cytoplasmatic MAP3K; infection; osmotic challenge in T.bruceiBaker et al., 2021 [58]
Agron et al., 2005 [57]
Fernandez-Cortes et al., 2017 [76]
CK1.2LdCK1.2Exosomal kinase; phosphorylates HSP90 and HSP23Hombach-Barrigah et al., 2018 [79]
Kröber-Boncardo et al., 2020 [87]
Other/CK2LmxCK2A1, LmxCK2A2T.brCK2 linked to cytoskeletal processes;
LbrCK2 secreted and ekto-forms mediating virulence
De Lima et al., 2006 [59]
Zylbersztejn et al., 2015 [74]
Dutra et al., 2009 [63]
PEKLmxEIF2αK2Vital for infection; T.brEIF2αK2 linked to sensing or transportBaker et al., 2021 [53]
Moraes et al., 2007 [61]
PIKK relatedLmxTOR3Infection; acidocalcisome formation and metabolic regulationBaker et al., 2021 [53]
Madeira da Silva and Beverley, 2010 [65]
CAMKLmxAKB1Infection; T.brAKB1: cytokinesis and divisionInoue et al., 2015 [73]
AGC/PKALmxPKAC3Infection; morphogenesisFischer Weinberger et al., under review [53]
Other/ULKLmxSTK36, LmxULK4Infection of sandfly vector and mammal. functionally linked. T.brSTK36, Tbr.ULK4: motility and flagella assemblyBaker et al., 2021 [51]
Varga et al., 2017 [76]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Clos, J.; Grünebast, J.; Holm, M. Promastigote-to-Amastigote Conversion in Leishmania spp.—A Molecular View. Pathogens 2022, 11, 1052. https://doi.org/10.3390/pathogens11091052

AMA Style

Clos J, Grünebast J, Holm M. Promastigote-to-Amastigote Conversion in Leishmania spp.—A Molecular View. Pathogens. 2022; 11(9):1052. https://doi.org/10.3390/pathogens11091052

Chicago/Turabian Style

Clos, Joachim, Janne Grünebast, and Myrine Holm. 2022. "Promastigote-to-Amastigote Conversion in Leishmania spp.—A Molecular View" Pathogens 11, no. 9: 1052. https://doi.org/10.3390/pathogens11091052

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop