Next Article in Journal
Sirtuin1-Mediated Deacetylation of Hypothalamic TTF-1 Contributes to the Energy Deficiency Response
Next Article in Special Issue
Investigating Cryptosporidium spp. Using Genomic, Proteomic and Transcriptomic Techniques: Current Progress and Future Directions
Previous Article in Journal
Emerging Roles of Lysophosphatidic Acid in Macrophages and Inflammatory Diseases
Previous Article in Special Issue
Trypanosoma cruzi Secreted Cyclophilin TcCyP19 as an Early Marker for Trypanocidal Treatment Efficiency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tackling Sleeping Sickness: Current and Promising Therapeutics and Treatment Strategies

by
Miebaka Jamabo
1,†,
Maduma Mahlalela
1,†,
Adrienne L. Edkins
2 and
Aileen Boshoff
1,*
1
Biotechnology Innovation Centre, Rhodes University, Makhanda 6139, South Africa
2
Department of Biochemistry and Microbiology, Biomedical Biotechnology Research Centre (BioBRU), Rhodes University, Makhanda 6139, South Africa
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(15), 12529; https://doi.org/10.3390/ijms241512529
Submission received: 13 June 2023 / Revised: 27 July 2023 / Accepted: 3 August 2023 / Published: 7 August 2023
(This article belongs to the Special Issue Modern Strategies for Diagnosis and Treatment of Parasitic Diseases)

Abstract

:
Human African trypanosomiasis is a neglected tropical disease caused by the extracellular protozoan parasite Trypanosoma brucei, and targeted for eradication by 2030. The COVID-19 pandemic contributed to the lengthening of the proposed time frame for eliminating human African trypanosomiasis as control programs were interrupted. Armed with extensive antigenic variation and the depletion of the B cell population during an infectious cycle, attempts to develop a vaccine have remained unachievable. With the absence of a vaccine, control of the disease has relied heavily on intensive screening measures and the use of drugs. The chemotherapeutics previously available for disease management were plagued by issues such as toxicity, resistance, and difficulty in administration. The approval of the latest and first oral drug, fexinidazole, is a major chemotherapeutic achievement for the treatment of human African trypanosomiasis in the past few decades. Timely and accurate diagnosis is essential for effective treatment, while poor compliance and resistance remain outstanding challenges. Drug discovery is on-going, and herein we review the recent advances in anti-trypanosomal drug discovery, including novel potential drug targets. The numerous challenges associated with disease eradication will also be addressed.

1. Introduction

African trypanosomiasis is a neuropathic and wasting disease, affecting both humans and animals, and endemic to Sub-Saharan Africa [1,2,3]. The causative agents of the disease in humans are subspecies of the extracellular vector-borne parasite Trypanosoma brucei, which is a protozoan parasite transmitted through the bite of an infected tsetse fly (Glossina species) during a bloodmeal [2]. The human infective forms, T. brucei gambiense (T. b. gambiense) and T. brucei rhodesiense (T. b. rhodesiense), are phenotypically indistinguishable and cause human African trypanosomiasis (HAT), commonly known as sleeping sickness, which is a neglected tropical disease (NTD) prevalent in Sub-Saharan Africa. T. b. gambiense is anthroponotic as humans are the main reservoir, while T. b. rhodesiense is zoonotic since its transmission relies on animal reservoirs [4]. The chronic infection caused by T. b. gambiense accounts for 98% of reported cases (referred to as G-HAT), while the acute zoonotic infection caused by T. b. rhodesiense is responsible for 2% of the reported cases (referred to as R-HAT) [5]. T. b. gambiense and T. b. rhodesiense are the only trypanosomes in Africa that have been reported to successfully establish an infection in humans [6]. T. brucei brucei (T. b. brucei) is one of several species of trypanosomes that cause nagana or Animal African Trypanosomiasis (AAT) in both domestic and wild animals [7].
All species and subspecies in the genus Glossina have the ability to transmit trypanosomes to humans, but some groups have been particularly adapted for transmission of specific species; the Glossina palpalis and Glossina fuscipes groups for T. b. gambiense and the Glossina morsitans and Glossina pallidipes groups for T. b. rhodesiense [8,9]. Although parasite transmission is primarily due to the bite of the tsetse fly, other modes of transmission have been reported, such as mother-to-child transmission through the placenta (for T. b. gambiense), mechanical transmission by other hematophagous insects, mishandling of laboratory samples, and sexual intercourse involving infected individuals [10,11,12].
It is difficult to determine the exact number of cases of the disease, as most infections now occur in remote rural foci. In April 2020, the World Health Organization (WHO) postponed all active screening campaigns for neglected tropical diseases; however, this recommendation was later revised to support NTD programs in the context of the COVID-19 pandemic. The impact of delayed screening in previous years had led to a resurgence of HAT in the endemic communities. The COVID-19 pandemic contributed to the lengthening of the proposed time frame of eliminating HAT as the control programs were interrupted and funding was diverted. It is predicted that the Democratic Republic of Congo (DRC) may experience the greatest risk due to several foci with high disease prevalence, coupled with violent conflict. If the goal to eliminate the transmission of HAT by 2030 is to be achieved, then urgent strategies need to be put in place to get treatment and screening campaigns back on track, to alleviate the losses experienced during COVID-19. Other aggravating factors include climate change, the increased movement of people and animals, and the urgent need for less complicated methods of diagnosis and treatment. Despite the recent development of a new drug, fexinidazole, novel anti-trypanosomal drugs are still being explored.

2. Disease Burden of Human African Trypanosomiasis

HAT is endemic to the Sub-Saharan African region, with rural farming communities being the most vulnerable [13,14]. The disease is distributed across 36 Sub-Saharan African countries, with T. b. gambiense present in 24 countries in Central and Western Africa, while T. b. rhodesiense is found in 13 countries in East and Southern Africa (Figure 1) [4,13,15,16]. Only in Uganda have both T. b. gambiense and T. b. rhodesiense been reported to occur, albeit in different regions: T. b. gambiense in the northwest and T. b. rhodesiense in the central regions (Figure 1) [17,18]. Knowledge of the tsetse fly involves their specific control methods, climate involvements, and other demographics that have significant roles to play in the prevalence of sleeping sickness in these areas [19]. The geographical prevalence of trypanosomiasis in Uganda, over the years, has shifted focus to the West Nile areas of Uganda [20,21].
HAT has caused millions of fatalities due to numerous outbreaks, which have led to epidemics. Over the course of the past two centuries, there have been three major epidemics in Africa. The first epidemic occurred between 1896 and 1906, killing nearly a million people [22,23]. East African countries such as Uganda and Kenya were severely affected by what was then assumed to be T. b. gambiense, as T. b. rhodesiense was yet to be characterized [24]. The second epidemic occurred about 10 years later and resulted in continuous surveillance strategies of the population, as well as vector control measures. These strategies were effective enough to almost eradicate the disease in less than 50 years [25]. However, a decline in surveillance measures and civil unrest in these countries led to a resurgence of the disease in numerous countries in Central Africa [26,27,28,29]. After the peak in the early 21st century, sustained efforts by agencies such as WHO to control G-HAT again included providing free drugs to affected areas, supporting and strengthening vector/disease control measures, and improving knowledge of the disease [30]. Consequently, the reported cases have dropped considerably over the years, from 40,000 cases in 1998 to 9878 cases in 2009. In 2019, the number of reported cases declined to 992, with over 70% of cases reported between 2009 and 2019 having been recorded in the DRC [31]. Despite the strides made in combating the disease, approximately 60 million people are still at risk of contracting HAT [2,32].
Despite the declining numbers of reported cases in the past decades, WHO indicates a wide discrepancy between the actual numbers of cases compared to the reported cases, due to poor coverage from inefficient surveillance systems [33]. There are reports of “blind spots”, where no HAT control activities have taken place in the past few decades in very remote areas of Central African Republic and the DRC. With areas such as these accounting for “invisible” cases, HAT elimination is complicated further. Incidental cases of HAT do get reported in non-endemic territories, even outside of Africa, which are attributed to infected individuals travelling from endemic countries [34]. Climate change is also expected to play an important role in the epidemiology of HAT, as global temperatures continue to rise [35].

3. Lifecycle of Trypanosoma brucei

T. brucei cycles between two obligatory hosts, the tsetse fly and the mammalian host, to complete its life cycle. Both the male and female blood-feeding tsetse fly can cause transmission [25]. The parasite morphotypes in the human and tsetse fly are referred to as the bloodstream (BSF) and procyclic (PCF) forms, respectively. The tsetse fly takes up the bloodstream trypomastigotes from the blood of the mammalian host while feeding, and the trypanosome multiples in the midgut, producing replicative procyclic trypomastigotes that enable the survival of the trypanosome in this environment (Figure 2) [36]. After complex adaptations through the tissues of the tsetse fly, parasites leave the midgut as epimastigotes and travel to the salivary glands where they multiply and form the short and stumpy metacyclic trypomastigotes, which are human infective. The parasite is then transmitted to a mammalian host from the bite of an infected fly (Figure 2) [37,38]. The BSFs of the parasite are either long slender (LS) or short stumpy (SS). The LS bloodstream forms are proliferative, adapted for optimal tissue invasion, and possessing the ability to traverse blood vessels and enter perivascular spaces [39,40]. During infection within the mammalian host, the LS BSFs differentiate into the quiescent SS BSFs. This is to ensure that the mammalian host lives long enough so that the parasites can be transmitted to another vector, which will in turn spread them to other mammalian hosts [4,41]. The transition from the LS morphology to the SS bloodstream forms is driven by a quorum-sensing pathway whereby high population densities of the LS BSFs result in the release of the stumpy-inducing factor [42]. The differentiation to the LS morphology is to ensure parasite survival, as the SS morphology is not optimally adapted for survival in the mammalian host for extended periods of time [43].
Generally, in a population of tsetse flies, the population carrying the infectious metacyclic form in their salivary glands is less than 0.1%, but with its feeding pattern within a 2–3 month lifespan, it can infect many people [38]. In the mammalian host, after the tsetse fly injects the metacyclic trypomastigote form before its bloodmeal, the trypanosomes first proliferate in the tissues at the site of the infection [39]. The parasites are transmitted as the tsetse fly injects its saliva in order to inhibit blood clotting and vasoconstriction [40,44,45,46].
One of the morphological markers is the position of the flagellar pocket relative to the nucleus [47,48]. T. brucei is restricted to the trypomastigote and epimastigote morphotypes, with the flagellar pocket of the trypomastigotes positioned at the posterior end of the cell opposed to the center in epimastigotes [2]. These morphotypes of T. brucei are characterized by having a laterally fixed flagellum [49].
T. brucei is an extracellular parasite generally shown to reside and multiply in the interstitial spaces, lymphatic system, and bloodstream of tissues of the mammalian host [4,50]. However, several other tissues have been found to serve as reservoirs for the T. brucei parasite, including the skin and adipose tissues [51,52,53]. An adipocyte tissue form, similar in morphology to the BSF, has been reported that invades mammalian fat tissues, and can utilize exogenous fatty acids, such as myristic acid, as a carbon source, and this may be a causative factor in the weight loss (wasting) that is observed in HAT sufferers [47]. Within the adipose tissues, the parasites occupy interstitial spaces, either between adjacent adipocytes or flanked by an adipocyte and a capillary [47]. The parasites possibly scavenge free fatty acids released by the adipocytes via an unknown mechanism [54]. T. brucei can carry out β-oxidation of fatty acids, but it is unknown if this pathway serves to fulfill the energy requirements of the adipose tissue forms [47]. A subpopulation of metacyclic trypomastigotes is also retained intradermally, within the vicinity of the bite, which remains highly infectious, even in the absence of detectable parasites in the blood [51]. In addition to the skin and adipose tissues, there is also evidence of the parasite residing in the testes, leading to sexual transmission, as shown in mice [55,56]. These reservoirs could be an extra layer of protection of the parasite from the host system, and a reason for continuous relapse in infected individuals. At the advanced stage of the parasitic infection, the parasites are also present in the cerebrospinal fluid, having infiltrated the central nervous system (CNS) [25].
As the parasite shuttles between the parasite vector and its mammalian host, it undergoes changes in its gene expression to provide proteins that are adapted to function in each host [57]. The BSFs uses glucose through the glycolytic pathway in the glycosome as they can barely survive in anoxic conditions, whereas the PCFs makes use of amino acids, such as proline and threonine, as carbon source through the Krebs cycle in the mitochondrion [58].

4. Symptoms and Disease Progression

The infected patient exhibits few symptoms immediately after being infected, but symptoms develop as the parasite multiplies in the blood and lymphatic vessels. The symptoms of HAT vary according to the stage of the parasitic infection (Figure 3) [2,59]. After the parasite transmission into the human bloodstream, headaches, fatigue, general malaise, and fever occur [13]. At the site of the infection, a lesion known as a trypanosomal chancre may also occur, particularly in cases of R-HAT [60]. As the infection progresses, the parasites invade lymph nodes, resulting in lymphadenopathy (Figure 3) [14]. The associated symptoms are heightened fever, chills, and hepatosplenomegaly [14,61]. At the phase of bloodstream and lymph node infection, the disease is said to be at the hemolymphatic stage [25]. In an advanced infection, the parasites breach the blood–brain barrier (BBB) to infiltrate the CNS, producing the meningoencephalitic stage (Figure 3) [25], which is characterized by mental, neurological, and sensory degeneration. The symptoms include, but are not limited to, delirium, disorientation, apathy, anxiety, emotional instability, abnormal speech, paresthesia, anesthesia, convulsions, seizures, and coma [13,61]. The meningoencephalitic phase of the disease is also symptomized by a disruption of the circadian cycle, resulting in daytime somnolence and nocturnal insomnia (Figure 3), hence the name sleeping sickness. If not treated, the meningoencephalitic stage of HAT can lead to coma and eventual death [4,59].
An early meningoencephalitic (“intermediate”) stage of the disease has also been suggested, whereby the parasitic infection has breached the BBB but has not yet infiltrated the brain parenchyma [62,63,64]. The rate of disease progression differs between R-HAT and G-HAT [32]. For R-HAT, the hemolymphatic stage symptoms appear 1 to 3 weeks following the bite of the tsetse fly, while the symptoms manifest much later in the case of G-HAT [13,60]. The onset of the meningoencephalitic stage of the disease occurs within 2–60 days of infection in cases of R-HAT, and death can take place within 3 months [14]. For G-HAT, the onset of the meningoencephalitic stage occurs between 300 and 500 days from infection, and the disease may persist for years [25,65].
In addition to the symptoms, especially in cases of R-HAT, cardiac and endocrine issues such as myocarditis and hypogonadism may also occur [66,67,68].
Figure 3. Progression of human African trypanosomiasis. The diagram highlights some of the defining symptoms of HAT as they relate to the progression of the T. brucei infection. The dashed line demarcates the point at which the parasitic infection infiltrates the CNS. The arrows at the bottom of the infographic indicate the different rates at which R-HAT and G-HAT progress. Adapted from [69].
Figure 3. Progression of human African trypanosomiasis. The diagram highlights some of the defining symptoms of HAT as they relate to the progression of the T. brucei infection. The dashed line demarcates the point at which the parasitic infection infiltrates the CNS. The arrows at the bottom of the infographic indicate the different rates at which R-HAT and G-HAT progress. Adapted from [69].
Ijms 24 12529 g003

5. Disease Diagnosis

Timely and accurate diagnosis is of paramount importance. Early diagnosis is key to ensuring that the patient receives treatment before the onset of the meningoencephalitic stage of the disease, while accuracy prevents a misdiagnosis, since some symptoms mirror those of malaria [64]. To prevent misdiagnosis, a microscopic analysis is carried out on the buffy coat of the blood sample [39]. Prior to microscopic analysis, the parasites may also be eluted from the blood sample by means of the mini anion exchange centrifugation technique (mAECT), in cases of low trypanosome concentrations [39,70,71]. The above-mentioned diagnostic techniques are not always useful for G-HAT diagnosis. Therefore, the card-agglutination trypanosomiasis test (CATT) is also carried out [72]. This diagnostic technique is specific to T. b. gambiense, detecting the presence of antibodies to VSG variants LiTat 1.3 and LiTat 1.5 [64,73]. Since treatment for HAT is infection-stage-specific, lumbar punctures are also administered as a supplementary technique, to determine if the infection has infiltrated the CNS [64]. In cases of low trypanosome concentrations in the blood, diagnosis may also be carried out using lymphatic fluids drawn from the swollen cervical lymph nodes of patients [39].

6. Anti-Trypanosomal Drug Treatments and Resistance Mechanisms

Even though African trypanosomes were first described more than a century ago, only a few efficacious drugs have been approved for treatment. These drugs present challenges that include toxicity, difficult administration, and trypanosome resistance. Administering the drugs is costly and labor intensive in some cases, a disadvantage considering that HAT is endemic to remote and resource-lacking regions [5,59]. At present, there is no umbrella treatment drug or strategy for HAT. The anti-trypanosomal chemotherapeutic treatment that is administered needs to be specific to whether the disease is at the hemolymphatic or meningoencaphalitic stage and whether the patient is infected with T. b. gambiense or T. b. rhodesiense [2,5,25].

6.1. Approved Drug Treatments

The approved drugs used to treat HAT have differing modes of action. The hemolymphatic stage of the disease is treated with pentamidine isethionate (pentamidine) and suramin for G-HAT and R-HAT, respectively. The meningoencaphalitic stage is treated with melarsoprol (Mel B or Arsobal®) and eflornithine (alpha-difluoromethylornithine, abbreviated to DFMO). DFMO is used in cases of meningoencephalitic G-HAT, while melarsoprol is used in cases of both G-HAT and R-HAT [2,5]. To expedite treatment of meningoencephalitic G-HAT, DFMO is also being used as a combinatorial drug in conjunction with nifurtimox (Lampit®) in what has been dubbed the nifurtimox–eflornithine combinatorial therapy (NECT). Nifurtimox is an independent drug that has not been authorized for the treatment of HAT, but is rather used for treating Chagas disease, that is caused by Trypanosoma cruzi, an American trypanosome [74,75,76,77].
The distinguishing factor between haemolymphatic and meningoencephalitic HAT treatment is whether the drugs are able to cross the BBB [78,79]. Pentamidine, which is indicated for hemolymphatic HAT, is BBB penetrative [80]. Pentamidine, which is administered through intramuscular injections, is an aromatic diamine, and its anti-trypanosomal activity dates back to the 1930s. Having been reported as effective against early meningoencephalitic HAT (“intermediate HAT”), the drug actively diffuses into the CNS in a process actively facilitated by the mammalian organic cation transporter 1 (OCT1) at the BBB [81,82]. This drug hinders DNA replication by influencing topological changes in the parasite’s DNA, thereby inhibiting normal topoisomerase functioning [2]. Pentamidine accumulation in trypanosomal cells also hinders mitochondrial activities [2,5]. The inadequacy of pentamidine as a trypanocide in the CNS is attributed to the drug accumulating at the capillary endothelium, and its active ejection back into the bloodstream by mammalian BBB ABC transporters, such as P-glycoprotein and multidrug resistance-associated proteins [80,82].
Suramin, dating back to the 1920s, is a polysulfonated naphthalene dye that inhibits glycosome-based glycolysis by interacting with selected enzymes that include 6-phosphogluconate dehydrogenase. Suramin anti-trypanocidal activity is also thought to be brought about by inhibiting low-density lipoprotein uptake, consequently having an adverse effect on the parasite’s cholesterol and phospholipid supply. Suramin is administered by means of intravenous injections. Even though suramin is only administered for treating R-HAT, it is also effective against G-HAT [2,5].
Melarsoprol, a derivative of melamine arsenical melarsen, has been used as a trypanocidal agent since the 1940s. The mode of action is not clearly understood, but it is hypothesized that the drug adversely modulates the parasite’s glycolytic and redox metabolism pathways. This drug is administered by means of an intravenous injection [5,83,84].
DFMO is a trypanocidal agent that was first used as a potential anti-cancer agent. Its use dates to the 1980s, and it acts by inhibiting ornithine decarboxylase (ODC) [85,86]. ODC is required for synthesizing polyamines that are essential for trypanosomal proliferation by cell division [86]. ODC is also necessary to produce molecules that serve as precursors of trypanothione, which is in turn necessary for the redox homeostasis of the parasite [87,88]. DFMO is administered through intravenous infusions [2,5,89]. The NECT strategy has been the main treatment of G-HAT since 2010. In addition to DFMO’s mode of action, nifurtimox places the trypanosomal cells under oxidative stress. This combinatorial therapy is administered orally and intravenously for nifurtimox and DFMO, respectively [5,90].

6.2. Drug Resistance Mechanisms

Several genotypic and molecular discoveries have been linked to observed drug resistance trypanosome phenotypes. Transporters are important for essential nutrient uptake in the T. brucei as parasites lack many anabolic pathways [91], such as the purine synthesis pathway, which is absent in protozoan parasites [92]. Some amino acids must be retrieved from outside of the parasites [93]. The vast majority of transporters are non-essential for parasitic survival due to redundancy, as a single substrate may enter through numerous transporters [91]. Since drug resistance is associated with transporter loss or mutation, it is hypothesized that the essential nutrients may still enter through other closely related or dissimilar transporters, which are non-permissive for the drugs [91]. Transporters that are responsible for drug resistance or treatment failure are non-essential for parasite survival; thus, receptor-mediated endocytosis is viewed as a drug target that could be manipulated for drug delivery [94].
The resistance of trypanosomes to pentamidine and melarsoprol is linked to a phenomenon referred to as melarsoprol/pentamidine cross resistance (MPXR) [95]. The implicated entities in MPXR are the aminopurine transporter P2 (encoded by TbAT1), the low-affinity pentamidine transporter (LAPT1), and the high-affinity pentamidine transporter (HAPT1; aquaglyceroporin2) [95]. With regard to the P2 transporter, it has been determined that its deletion or loss only partially influences MPXR, whereby HAPT1 loss is the primary determinant for high-level MPXR [95,96,97,98]. LAPT1 has a significantly lower affinity for pentamidine compared to the P2 transporter and HAPT1 [99,100]. HAPT1 has the highest affinity for pentamidine, while the P2 transporter is the primary factor in melarsoprol uptake [99,100].
There are two orthologues of HAPT1 in T. brucei, aquaglyceroporins 1 and 3 (AQP1 and AQP3) [98,101]. Field studies have shown that T. brucei mutants exhibiting AQP2 and AQP3 chimerization are responsible for pentamidine and melarosprol resistance, even in the absence of mutations or losses of the P2 transporter [102,103,104]. AQP3 is inconsequential in terms of MPXR [96]. AQP3 is localized at the cell body membrane and AQP1 at the flagellar membrane [96,101]. AQP2 localizes at the flagellar pocket and flagellar membrane in the BSFs and PCFs, respectively [96]. Since endo- and exocytosis are increased at the flagellar pocket, it is possible that the localization of AQP2 could be a factor in the transporter being the main determinant of pentamidine and melarsoprol uptake. The ability of AQP2 to transport the drugs is attributed to its enlarged atypical core, which is markedly different from those of AQP1 and AQP3 [96,100,105]. There seems to be contention when it comes to the pentamidine uptake mechanism by AQP2. Whether the drug permeates through the core of AQP2 or AQP2 serves as a receptor for the endocytic uptake of the drug is unclear [105,106].
Coupling pentamidine to PEGylated chitosan particles coated with a camel heavy-chain-derived nanobody resulted in increased drug uptake, treating HAT in mice models at 100-fold lower concentrations compared to pentamidine alone [107]. This is due to the nanobody recognizing cryptic epitopes that are presented at the trypanosomal cell surface [107,108]. The pentamidine loaded onto the PEGylated chitosan particles coated with the nanobodies (NbAn33-pentamidine-chNPs) was also trypanocidal in vitro toward resistant AQP2-lacking cell lines [107]. NbAn33-pnetamidine-chNP was also able to treat mice infected with an AQP2-lacking trypanosomal cell line [107]. AQP2-lacking trypanosomes are highly sensitive to salicylhydroxamic acid (SHAM), octyl gallate, and propyl gallate, which are all inhibitors of the trypanosome alternative oxidase (TAO) [109]. This is due to AQP2 being responsible for glycerol uptake and efflux; therefore, inhibiting TAO leads to pathways that result in a toxic accumulation of glycerol [91,109].
Surface glycoproteins have been implicated in suramin resistance. Suramin interacts with invariant surface glycoprotein 75 (ISG75), where the abundance of ISG75 is a function of suramin accumulation in the trypanosomal cells [110]. Cells displaying a particular variant surface glycoprotein (VSG), denoted as VSGsur, exhibited an increased resistance toward suramin [111]. A high-affinity complex is formed between suramin and VSGsur, and the drug is bound by the large pocket of the VSG homodimer [112]. Mutations within the suramin-binding pocket of VSGsur perturb the high-affinity VSGsur–suramin complex [112]. The mechanisms of suramin resistance by VSGsur is that the VSG impairs particular receptor-mediated endocytic pathways that are used by trypanosomes to internalize suramin [113].
VSGsur is a member of a group of VSGs that exhibit unique structural properties [112], and within this group, VSG13 has been identified and structurally defined [112]. It has not been determined if VSG13 and other VSGs that fall within this structural grouping also possess suramin-binding capabilities or if they serve as determinants for suramin resistance when presented at the trypanosomal cell surface. A study of trypanosomes obtained from suramin responsive and non-responsive HAT patients attempted to elucidate the underlying proteomic profiles of resistant trypanosomes [114]. The trypanosomes isolated from the non-responsive patients (T. b. rhodesiense, strain EATRO-734) exhibited upregulated metabolic and detoxification processes [114]. Furthermore, the mitochondrial proteome of the EATRO-734 strain was also demonstrated to be upregulated. Inadequacies in the endosomal pathway are also implicated in suramin resistance [114]. Regarding DFMO, a single gene (TbAAT6) has been determined to encode the transporter of the drug, whereby the loss or knockdown of the gene results in drug resistance [115]. It is suggested that DFMO resistance may easily be detected in the field by carrying out PCR [115].

6.3. Physiological Challenges to Effective Drug Delivery

An increased understanding of trypanosomal biology is required to facilitate effective drug delivery as the parasite resides in different tissue types in the body, effectively serving as a long-term reservoir. Due to renewed interest in tissue-resident parasite populations, the implications of these parasite populations for future studies were reviewed [116]. During the course of infection, subpopulations of the parasites may exhibit divergent metabolic profiles where some subpopulations are quiescent, with downregulated metabolic pathways that are targeted by drugs [117]. Some subpopulations could be at different stages of differentiation; therefore, drugs targeting cell division may be ineffective. These subpopulations, which may be less susceptible to being targeted by drugs, are referred to as persisters [117,118], and may cause relapses in infectious diseases, even after treatment courses have been duly completed [118,119].
In the protozoan parasites Plasmodium vivax and Toxoplasma gondii, persister subpopulations, represented by the hypnozoite and bradyzoite morphotypes, respectively, have been reported [118,120]. In Chagas disease, the persister phenomenon is implicated in the chronicity of the disease and the frequency of non-efficacy by the front-line drugs, nifurtimox and benznidazole [121,122]. Eliminating the parasitic population in its entirety is essential for curing Chagas disease, as the persistence of low populations of the parasite could worsen the outcomes for patients by eventually resulting in sickness in asymptomatic individuals [117].
Tissue distribution also needs to be considered as African trypanosomes are dispersed across several tissues, with some tissues potentially offering protection from the immune system [123]. This is because immune-privileged and drug-impenetrable tissues may be sources of infection re-emergence, even in cases where the anti-parasitic treatment has been administered to the full term [123]. The T. brucei parasite in the early stage of the infection invades various sites, which apart from the bloodstream and lymph nodes, are the testes, skin, and adipose tissue [47,52,55]. These sites are believed to be immune privileged, possibly providing a safe haven for parasitic cells [123]. However, there are emerging reports of trypanosomes accumulating in adipose tissue, whereby they trigger an immune response [124]. The immune response leads to adipocyte lysis (adipolysis) [124]. In this instance, adipolysis is hypothesized to serve as a parasitemia reduction mechanism, also prolonging the lifespan of the mammalian host [47,125]. The adipose tissue forms of the parasite are quiescent, exhibiting downregulated protein synthesis, and they are less susceptible to drugs [124]. They exhibit reduced virulence and eventually become undetectable, being a possible source of disease re-emergence. The adipose tissue forms are therefore a potential source of persistence and chronicity [124]. A study found that persister quiescent skin tissue forms of the T. brucei parasite, which are hypothesized to be responsible for prolonged and persistent infections, were quickly established when cultured in artificial human skin, and these quiescent forms could be metabolically reactivated [126]. Therefore, the drug impenetrability of adipose tissues presents an additional challenge [47,123]. The adipose-tissue-occupying parasites are possibly more densely populated compared to those in the bloodstream and CNS [126]. Adipose- and skin-tissue-occupying parasites could be targeted via the fatty acid uptake machinery of the parasite, as it is viewed as a vehicle for drug delivery [54]. This entails coupling drugs to fatty acids, which would be lipophilic and thus optimized for skin or even adipocyte tissue penetration [54].

6.4. Recent Advances in Drug Development

Most recently, fexinidazole has been sanctioned for the treatment of both stages of G-HAT, albeit indicated for non-severe meningoencephalitic HAT [127,128,129]. Fexinidazole is a pro-drug that emerged from a Drugs for Neglected Diseases initiative (DNDi) and Swiss Tropical and Public Health Institute (Swiss TPH) collaboration, and was further developed by Sanofi. The drug is currently the only HAT treatment that is administered orally, having efficacy at both stages of the disease [129]. Fexinidazole, like nifurtimox, is a nitro pro-drug dependent on the putative ubiquinone nitroreductase for activation in the mitochondrion. It is however unclear whether the mode of trypanocidal action is mainly due to mitochondrial function inhibition or the targeting of components outside the parasite’s mitochondrion [130,131,132,133]. Fexinidazole is being trialed for the treatment of both stages of R-HAT and intermediate Chagas disease. The R-HAT clinical trials were conducted in Uganda and Malawi at the Lwala and Rumphi district hospitals, respectively [134]. Uganda and Malawi account for more than 90% of R-HAT incidences. The outcomes of the fexinidazole trials against R-HAT are currently being prepared for publication [134]. The phase II, multicenter, randomized, and placebo controlled proof-of-concept trial of fexinidazole against intermediate Chagas has recently been published [135]. Fexinidazole was determined to effectively clear trypanosomes in Chagas patients; however, cases of neutropenia and upregulation of hepatic enzymes were also detected [135].
In addition to fexinidazole, which has been approved for clinical use, several drugs with the potential to yield the effective treatment of HAT have recently progressed to clinical trials [136]. A class of compounds, referred to as benzoxaboroles, has emerged as promising novel drugs for the treatment of HAT [137]. The compounds are derivatives of an oxaborole heterocycle (boron heterocyclic) that is coupled to a phenyl group [138,139]. The most notable benzoxaborole is acoziborole (AN5568, SCYX-7158), which is derived from gem-dimethyl 4-fluoro-2-trimethylfluoro benzamide that has progressed to phase III clinical trials for G-HAT, and it is leading in the pipeline of novel HAT drugs [136,140]. Orally dosed acoziborole can efficaciously and safely treat the hemolymphatic and meningoencephalitic stages of G- and R-HAT [141]. SCYX-7158 is metabolized into SCYX-3109 by trypanosomatids, with its cellular target being the nuclear-localized cleavage and polyadenylation specific factor 3 (CPSF3), an mRNA processor [139,141]. CPSF3 forms part of the pre-mRNA cleavage and polyadenylation complex that cleaves the pre-mRNA before coupling the poly (A) tail [142,143]. CPSF3 is also involved in the histone cleavage complex that cleaves and processes pre-mRNAs of core histones without any adenylation taking place [142]. CPSF3 is also targeted by AN11736, another benzoxaborole, which is being developed as an anti-trypanosomal for the treatment of AAT caused by T. vivax and T. congolense [144]. The molecular docking of acoziborole to T. brucei CPSF3 reveals 26 highly conserved amino acids in close proximity to acoziborole [139]. Between human and T. brucei, there are only four amino acid residue substitutions at the sites that are in close proximity to docked acoziborole [139]. Therefore, the scientific explanation for the selectivity of acoziborole could be that the substitutions could decrease the drug’s affinity for the human orthologue of CPSF3.
Pafuramidine (DB289), a prodrug that is metabolized to form the active anti-trypanosomal diamidine analogue of pentamidine, furamidine (DB75), had also reached clinical trials [5,145,146]. Pafuramidine was the first orally administered drug to reach phase III clinical trials for the treatment of hemolymphatic HAT [146]. DB75, like pentamidine, is a diamidine that selectively accumulates in trypanosomes, with the DNA-containing organelles taking precedence [147]. The drug with an unknown mechanism of action was discontinued due to safety and toxicity concerns [146].
Drawing from pafuramidine, a series of other diamidine molecules have been investigated due to their anti-trypanosomal properties [5]. DB829 and DB820, which are the active metabolites derived from the prodrugs DB868 and DB844, respectively, have been determined to be efficacious against stage II HAT in animal models [148,149]. DB820 has also been determined to possess a similar mechanism of accumulation as pentamidine, accumulating in DNA-containing organelles and binding to A-T-rich sites on DNA [148].
DB829 is BBB penetrant, and an effective trypanocide within the CNS [149]. Due to its structural similarity to DB75, the development of DB829 was discontinued, but the structural differences may have rendered DB829 less toxic [150]. Since pentamidine and the other diamidines are ineffective against meningoencephalitic HAT, the efficacy of DB829 could be due to BBB transporters [150,151]. DB829 is either more effectively imported into the CNS by the transporters or the transporters that extrude pentamidine from the CNS are less efficacious against DB829 [80,150].
Due to the toxicity of melarsoprol and the inconvenience associated with its administration, melarsoprol–cyclodextrin complexes were developed as potential HAT chemotherapeutics [150,152]. The complexes, melarsoprol–hydroxypropyl-β-cyclodextrin and melarsoprol-randomly methylated-β-cyclodextrin, were determined to be efficacious against meningoencephalitic R-HAT when orally administered in mice [152]. This was viewed as a promising outcome and was given orphan drug status by the Food and Drug Administration (FDA) in the U.S., as well as the European Medicines Agency (EMA) [150]. Protocols for a phase II clinical trial of the melarsoprol–cyclodextrin complexes were developed; however, they never went ahead due to a lack of funding [150]. Successfully consolidating funds for studying arsenicals in humans is highly unlikely [150]. Cyclodextrins are known to improve the solubility of drugs without altering their physical properties [153].

6.5. Validated and Potential Drug Targets

A significant number of compounds exhibiting diverse molecular structures have shown antiparasitic effectiveness in the laboratory and are interesting lead compounds for the development of new drugs. The major validated drug targets in T. brucei are discussed, including other potential drug targets.
In the studies carried out on the various potential HAT drugs and their validated cellular targets, different subspecies and strains have been used, including the animal-infective T. b. brucei. Though phenotypically identical with T. b. rhodesiense, T. b. gambiense has been determined to be genetically divergent [154]. The T. b. brucei TREU927 reference strain exhibits all the known phenotypes of T. brucei organisms [155]. Therefore, discoveries made in studies where T. b. brucei is employed could be inferred for T. b. rhodesiense.

6.5.1. N-Myristoyltransferase

A selective anti-trypanosomal drug target with orthologues in humans is N-myristoyltranferase (NMT). NMT uses myristoyl-coenzyme A (CoA) as a substrate and functions in a co- or post-translational modification process referred to as myristoylation, whereby a myristate from myristoyl-CoA is transferred to an N-terminal glycine residue [156,157,158,159]. Myristoylation promotes the interaction of proteins with membranes [157]. The knockdown of NMT arrests parasitic growth and adversely affects infectivity [160,161]. Though ubiquitous and essential in eukaryotic life, it has been established that NMT inhibition in trypanosomes is highly deleterious. The difference between NMTs of humans and T. brucei is that humans have two genes encoding the protein, human NMT1 and NMT2, sharing 55% and 69% similarity with T. brucei NMT, respectively [162,163]. In T. brucei, phosphatases, calpain-like proteins, the ADP-ribosylation factor (ARF), and ARF-like families of GTPases are subject to myristoylation by NMT [164,165,166]. In L. major, hydrophilic acylated surface proteins undergo NMT-facilitated myristoylation, together with the flagellar calcium-binding protein (FCaBP) in T. cruzi [167,168]. Approximately 60 proteins are subject to N-myristoylation in kinetoplastids [169].
NMT inhibition is trypanocidal in vitro and in rodent models of T. brucei infections [170]. Classes of NMT inhibitors, referred to as sulfonamides, represented by DDD85646 have been investigated as potential anti-parasitic drugs [163]. DDD85646 was effective in treating mice infected with T. b. brucei S427 and fully curative in mice infected with T. b. rhodesiense STIB900 [170]. However, due to the conservation of NMTs between protozoan parasites and humans, there is a challenge of non-specificity or non-selectivity by the compounds as they are also potent against human NMTs [171]. Moreover, the BBB is impermeable for DDD85646, making the drug ineffective for meningoencephalitic HAT [163]. Therefore, optimizing candidate drugs to enable CNS penetration is also of interest, with a lead study having been conducted based on DDD85646 [172]. The DDD85646 optimization derivatives were demonstrated to potentially cross the BBB, possessing anti-trypanosomal properties [172]. The DDD85646 predicted binding site in T. brucei NMT, according to molecular docking, it comprises 31 amino acid residues, and human NMT1 and NMT2 share 83% and 90% similarity, respectively [163].

6.5.2. Ubiquitination/Proteasome

The trypanosomal protein ubiquitination and proteasomal degradation processes are also viewed as potential targets for HAT drugs [173,174]. Ubiquitination is a post-translational modification process that targets terminally damaged or aged proteins for degradation in the proteasome. Ubiquitination primarily targets lysine residues on the substrate protein, modifying the amino acid by attaching ubiquitin [175]. Ubiquitination is either mono or poly, whereby mono-ubiquitination refers to the ubiquitination of the substrate protein, and poly-ubiquitination entails the formation of ubiquitin chains whereby the lysine residues on ubiquitin are also ubiquitinated [176,177,178]. Ubiquitination is carried out by three classes of enzymes, E1 (ubiquitin-activating enzyme), E2 (ubiquitin-conjugating enzyme), and E3 (ubiquitin ligase), which act on the substrate protein consecutively, cascading from E1 down to E3 [175,179,180]. Polyubiquitination is generally accepted as a prerequisite for proteasomal degradation, and it is carried out by E2 and E3 ligases [181,182]. Mammalian cells only possess a single E1 (also referred to as ubiquitin-activating enzyme 1; UBA1), while kinetoplastids possess two enzymes. The T. brucei E1 proteins, TbUBA1a and TbUBA1b, share 36% and 24% sequence identity with the human UBA1, respectively [173]. TbUBA1b knockdown is highly deleterious, and it results in an arrest of ubiquitination [183,184]. There are estimated to be fifteen E2 and sixty E3 ubiquitin ligase enzymes in T. brucei [174,185].
In the protein ubiquitination–proteasome pathway in T. brucei, proteins that fail to be imported into the mitochondrion are targeted for proteasomal degradation [186]. Abrogating the ATOM69 component of the atypical mitochondrial outer membrane translocase (ATOM) complex results in the channeling of three proteins to the mitochondrion, all of which are trypanosomatid-specific and involved in the ubiquitination–proteasomal pathway [186]. The proteins are T. brucei ubiquitin-like domain (TbUbL1), T. brucei C-terminal HECT E3 ubiquitin ligase domain (TbE3HECT1), and a hypothetical protein [186]. TbE3HECT1 is an E3 ubiquitin ligase [186]. These three proteins are involved in the channeling of ATOM69-specific mitochondrial substrate proteins for degradation, preventing an accumulation of the mitochondrial proteins in the cytosol [186]. This trypanosomal mechanism places further importance on targeting the ubiquitination–proteasome pathway in trypanosomes.
Targeting the ubiquitination process at the E1 level is preferred as it is at the initiation of the cascade [187]. The trypanosomal proteasome is currently the subject of research into two orally administered azabenzoxazole compounds: GNF6702, which is an improved derivative of GNF5343, and GSK3494254 (or DDD01305143) [188,189]. The compounds have been shown to be efficacious against the viability of these three kinetoplastid pathogens, namely, Trypanosoma brucei, Trypanosoma cruzi, and Leishmania major (TriTryps). GNF6702 and GSK3494245 bind between the β4 and β5 subunits of the proteasome, inhibiting the chymotrypsin-like activity of subunit β5 [188,189]. There are solubility concerns with regard to orally administered GNF6702, which led to the molecule being coupled with a pyridine to yield the more soluble LXE408 [190]. The resultant LXE408 compound is also efficacious against Leishmania tarentolae [190]. The phase II clinical trials of LXE408 are scheduled to commence, having passed the phase I clinical trials for the treatment of PKDL [191]. GSK3494245 has also reached phase I clinical trials for visceral leishmaniasis [189].

6.5.3. Cyclic Adenosine Monophosphate-Specific Phosphodiesterases B1 and B2

In T. brucei, the B family of phosphodiesterases (PDEB) primarily functions in the hydrolysis of cellular cyclic adenosine monophosphate (cAMP) as a negative feedback mechanism, being essential for the BSF of the parasite [192]. The PDEB family consists of two paralogs (PDEB1 and PDEB2) that share 75% similarity and are both located at the flagellar membrane [192,193]. However, PDEB2 predominantly localizes in the cytosol [192]. The simultaneous knockdown of PDEB1 and PDEB2 increased intracellular cAMP levels, which was lethal for trypanosomes [192]. Morphological defects associated with PDEB1 and PDEB2 knockdown and inhibition are multinucleation and multiflagellation [194]. This has led to the phosphodiesterases being identified as drug targets in trypanosomes. Moreover, human phosphodiesterase inhibitors are already on the market for the treatment of ailments, such as chronic obstructive pulmonary disease, cardiovascular diseases and inflammation [195]. Phosphodiesterases are highly conserved between humans and the African trypanosome; therefore, the expertise in treating human diseases serves as a foundation for targeting these enzymes as drug targets from the parasite [194]. A known inhibitor of the T. b. brucei PDEBs is CpdA (now referred to as NPD-001), a potent tetrahydrophthalazinone compound that results in inhibition in the nanomolar concentration range [194]. Given that the PDEBs localize in the flagellar membrane, NPD-001 is lipophilic, which potentially eliminates resistance and uptake issues [196]. NPD-001 results in an increase in intracellular cAMP, leading to parasite mortality within 3 days [196]. The sensitivity of T. b. brucei to NDP-001 was linked to the cAMP response proteins (CARP), of which four paralogs were implicated (CARP 1–4) [196]. A further seven CARP genes (CARP 5–11) were also identified as inducing resistance to NPD-001 in T. b. brucei [197]. CARP 1 and CARP 11 have an affinity to NPD-001 [197]. CARP genes are more pronounced in kinetoplastids, with CARP 3 and CARP 11 being restricted to the Trypanosoma genus [197,198,199]. NPD-001 is non-specific, also being inhibitory toward human PDEB4. PDEB4 inhibition leads to the suppression of the tumor necrosis factor α (TNF- α) cytokine in humans [200,201]. As such, the development of the compound as a potential anti-trypanosomal drug in humans is unlikely.
A recent study has identified the anti-trypanosomal activity of phenylpyridazinone analogs of NPD-001 that inhibit T. brucei PDEB1 [202]. Furthermore, a selectively inhibitory alkynamide phthalazinone molecule has also been shown to be trypanocidal, having no toxic effects toward human MRC-5 cells [203]. Despite PDEB1 being a promising drug target, there appear to be no studies to determine the effects of PDEB1 inhibitors in animal models of T. brucei infection.

6.5.4. Oxidative Stress/Polyamine Synthesis—Trypanothione System

Trypanothione reductase (TyrR) is analogous to mammalian glutathione reductase, but these proteins share less than 50% sequence similarity [204,205]. TyrR is a flavoprotein that facilitates the reduction of trypanothione disulphide to produce a dithiol, which is functionally equivalent to glutathione in mammals and functions as an antioxidant [204,206,207]. Due to the considerable structural variation between TyrR and glutathione reductase, this trypanosomatid system presents a promising target in terms of drug discovery [205,208]. However, a major challenge with developing TyrR inhibitors is the enzyme’s enlarged hydrophobic binding site, which is not compact enough for high-affinity interactions with small-molecule inhibitors [209,210]. Also, to achieve an adequate trypanocidal effect, TyrR activity must be inhibited by up to 90% due to the enzyme’s elevated efficacy in terms of turnover [207,211]. A spiro derivative molecule has been determined to differentially inhibit trypanothione reductase, having no modulatory effect on human glutathione reductase [212]. Spiro molecules serve as a scaffold for BBB-penetrant molecules [212].
The trypanothione molecule itself is synthesized in a process facilitated by trypanothione synthetase (TyrS), which is also a validated drug target that has been reported to be essential for trypanosome survival [213,214,215]. TyrS forms part of the machinery that attaches two glutathione molecules to the polyamine spermidine to result in the formation of trypanothione [213,214]. Engineered strains of T. cruzi expressing elevated amounts of TyrS exhibit increased growth rates and tolerance toward oxidants and heavy metals [216]. Moreover, the strain also exhibited resistance toward Chagas drugs, beznidazole and nifurtimox [216]. Recent research has identified Ebselen as a tightly binding irreversible inhibitor of TyrS in T. brucei, having a trypanocidal effect on cultured T. b. brucei cells [217].
With regard to trypanothione production, DFMO may also hamper the molecule’s biogenesis [218]. Therefore, a TyrR inhibitor could be used in combination with DFMO in treating HAT, as with nifurtimox in the NECT [218,219]. This is probably due to DFMO inhibiting the polyamine synthesis pathway, which leads to the synthesis of spermidine, a precursor for trypanothione synthesis [218,220]. The decarboxylation of L-ornithine by ODC is rate-limiting, and it is the initial step in the synthesis of the polyamines putrescine, spermidine, and spermine [221]. Using putrescine as a substrate, spermidine synthase forms spermidine, which in turn serves as the precursor for the synthesis of spermine by spermine synthase [220]. Spermidine synthase is also validated as a drug target as it is essential for parasite survival [222]. The activities of spermidine and spermine synthases rely on the donation of the aminopropyl group by decarboxylated S-adenosylmethionine to their precursor substrates, putrescine and spermidine, respectively [220,222,223]. The decarboxylated S-adenosyl methionine is synthesized by S-adenosylmethionine decarboxylase (AdoMetDC), which is a validated drug target [221]. The activity of AdoMetDc is also rate-limiting in the polyamine synthesis pathway [221].
Therefore, the polyamine synthesis pathway in which ODC is involved could further be perturbed by inhibiting AdoMetDC, which acts downstream of ODC. A study has shown the potential of inhibiting AdoMetDC, whereby thirteen classes of compounds were identified [224]. Of the thirteen, eight compounds were shown to differentially inhibit AdoMetDC in T. brucei compared to the human orthologue, with some of the selective compounds potentially being CNS penetrative [224]. In mammals, AdoMetDC is an inactive proenzyme that is activated by serinolysis, which leads to the formation of β- and α-chains from the N- and C-terminal of the protein, respectively [225,226,227]. The β- and α-chains form protomers that oligomerize to form homodimers [227]. The attractiveness of inhibiting AdoMetDC arises from the fact that the enzyme is divergent in T. brucei compared to other eukaryotic organisms [228]. The AdoMetDC gene is duplicated in T. brucei to form AdoMetDC and an inactive (“dead”) peptide that is referred to as prozyme, whereby the duplicate gene products form a heterodimer [228]. The prozyme is an allosteric regulator, stimulating the activity of AdoMetDC by over a thousand-fold [228]. The prozyme interacts with a region of the N-terminus of the T. brucei AdoMetDC that is lacking in other eukaryotic organisms [229].

6.5.5. RNA-Editing Ligase and Pteridine Ligase

RNA-editing ligase (REL1) and pteridine reductase 1 (PTR1), which are unique to trypanosomatids, have also been validated as drug targets. REL1 in T. brucei is essential for survival in the tsetse vector and the mammalian host [230,231]. The ligase is one of two enzymes that function downstream of one of the trypanosomatid RNA-editing mechanisms, joining together the maxi-circle transcript (mRNA) molecule after the insertion of the deletion of uridine [230,231]. A study has also been conducted whereby potential REL1 lead inhibitors were identified [232]. PTR1 plays a secondary role in the processing of folates in trypanosomatids, with dihydrofolate reductase (DHFR) and thymidylate synthase (TS) playing a primary role [233,234]. The knockdown of PTR1 leads to major morphological deficiencies and is trypanocidal [234]. PTR1 knockdown trypanosomes also exhibit reduced pathogenicity [234]. DHFR-TS inhibitors are thought to be viable drug targets for the treatment of HAT; however, the inhibition of DHFR-TS results in the upregulation of PTR1, which compensates for the role of DHFR-TS [235]. Therefore, a chemotherapeutic strategy against both PTR1 and DHFR-TS could result in a potent anti-trypanosomal treatment [236,237]. Recent studies have also identified novel trypanocidal PTR1 inhibitors when combined with DHFR inhibitors. In combination with the DHFR inhibitor WR99210, a series of molecular docking predicted inhibitors of PTR1 were shown to inhibit the growth of T. b. brucei Lister 427 [238]. Furthermore, the inhibitors were also potential DHFR inhibitors [238]. In a separate study, cycloguanil, which is a DHFR inhibitor, has been determined to inhibit PTR1. Derivatives of cycloguanil were also demonstrated to possess EC50 values in the nanomolar concentration range against T. b. brucei Lister 427 bloodstream forms [239]. Though PTR1 has been reported to be essential for virulence in mice, no studies have been carried out to determine the effects of PTR1 and DHFR-TS inhibitors in animal models of trypanosomal infection [234].

6.5.6. Repurposing Anti-Cancer Drugs as Anti-Parasitic Treatments

The repositioning of anti-cancer agents as HAT drugs is premised upon the fact that trypanosomes rapidly proliferate within the mammalian host in a manner similar to cancer cells. Therefore, compounds that inhibit cell replication, such as DNA replication inhibitors, may have some anti-trypanosomal effects [86]. Proof of this concept is provided by pentamidine, which is an existing HAT drug that inhibits trypanosomal DNA replication [2]. Interestingly, recent research has confirmed the anti-cancer capabilities of pentamidine, even though the mechanism of action in this regard may not include DNA replication inhibition [240,241,242].
A number of glycolysis inhibitors with anti-cancer potential are active against T. brucei [243]. Protozoan parasites, like many cancer cells, have an upregulated glycolytic process [243]. Dichloroacetic acid (DCA), 3-bromopyruvic acid (3BP), lonidamine (LND), metformin (MET), and sirolimus (SIR) have all been determined to reduce the survival of T. brucei parasites [243]. Recently, a potent, novel anti-cancer agent MitoTam, a drug molecule resulting from the conjugation of a tamoxifen derivative to a triphenylphosphonium vector (TPP+), has been an efficacious trypanocidal against T. b. brucei at nanomolar concentrations [244,245]. MitoTam acts by disrupting the mitochondrial integrity of trypanosomes and has been determined to be minimally toxic against healthy mammalian cells [245]. MitoTam treatment was also shown to prolong the lifespan of T. brucei-infected mice, which when left untreated, die within eight days [245].

6.5.7. Heat Shock Protein 70 and 90

The T. brucei molecular chaperone network is implicated in survival, differentiation, and pathogenicity, as well as coping with environmental stressors [246,247,248,249]. An analysis of the molecular chaperone machinery in T. brucei revealed an expansion in the number of J-proteins and Hsp70 proteins, indicating that these protein families may play a critical role in the biology of the parasite [250]. Co-chaperones of T. brucei Hsp90 were recently identified, and it is possible that chaperone/co-chaperone interactions could be pursued as potential drug targets [251]. Molecular chaperones, in particular the Hsp70 and Hsp90 families, have been assessed as potential anti-trypanosomal drug targets [252,253,254].

Trypanosoma brucei Hsp70

The T. brucei cytosolic Hsp70 complement consists of three paralogues, namely TbHsp70 and TbHsp70.c, which are heat inducible, and TbHsp70.4, which is non-inducible [246,253,254]. TbHsp70 and TbHsp70.4 were susceptible to malonganenone and nuttingin compounds, which modulated their substrate-binding capabilities, and were trypanocidal [255]. As Hsp70s are ATP-dependent, they rely on type-I or type-II J-proteins to stimulate their ATPase activity. In this regard, the compounds inhibited the TbHsp70 and TbHsp70.4 ATPase activity stimulated by the cytosolic T. brucei type I J-protein, Tbj2 [255]. However, this inhibition was reduced in the J-protein-stimulated ATPase activity of the constitutively expressed human cytosolic Hsp70 [255].
Methylene blue and quercetin, which are trypanocidal, were reported to inhibit the activity of TbHsp70.c [256]. TbHsp70.c is an atypical Hsp70, possessing orthologues only within the order Trypanosomatida, making it a possible drug target [256]. TbHsp70.c possesses a divergent linker region, and substitutions at otherwise highly conserved amino acid residues within the substrate-binding pocket [256]. Furthermore, TbHsp70.c lacks the EEVD motif, which is essential for interacting with TPR domain-containing co-chaperones, such as the stress-inducible phosphoprotein 1 (STI1) [250,254]. STI1 serves as a regulator of substrate channeling between Hsp70 and Hsp90, and TbHsp70.c does not interact with T. brucei STI1 [256,257]. The benefit of targeting Hsp70s is that they possess deep nucleotide-binding clefts and substrate-binding pockets that can be targeted with small molecules [258]. The uniqueness of TbHsp70.c in this regard could serve as a basis for the development of drugs that differentially kill trypanosomal cells.

Trypanosoma brucei Hsp90

Hsp90 isoforms are important for the survival of cancer cells, and numerous anti-Hsp90 compounds have been identified and are in the pipeline for development as anti-cancer therapies. Indeed, the Hsp90α/β inhibitor, Pimitespib, received its first approval in Japan for the treatment of progressive gastrointestinal stromal tumor (GIST) [259,260,261,262]. Some of these compounds also possess potent anti-trypanosomal properties, showing an increased binding affinity for T. brucei Hsp90 compared to human Hsp90 [263]. The inhibition of cytosolic T. brucei Hsp90, TbHsp83, is anti-trypanosomal, whereby TbHsp83 is more sensitive to Hsp90 inhibitors compared to mammalian orthologues. T. brucei showed about 1000-fold higher sensitivity to geldanamycin and over 50-fold higher sensitivity to the geldanamycin analog 17-AAG (17-allylamino-17-demethoxy-geldanamycin/Tanespimycin) compared to its mammalian host. Another geldanamycin analog, 17-DMAG (17-dimethylaminoethylamino- 17-demethoxygeldanamycin), cured mice of T. brucei infection [264,265]. TbHsp83 is essential for the survival of the parasite [249,264]. The Hsp90 inhibitors, geldamycin and radicicol, had anti-trypanosomal EC50 that were on par with the existing HAT drugs, DFMO, suramin, melarsoprol, and pentamidine [265]. T. b. brucei MiTat 1.2 strain 427 bloodstream forms were used for the study [265]. The geldanamycin derivatives, 17-AAG and 17-DMAG, also selectively inhibited trypanosomal growth at nanomolar concentrations [265]. Mammalian cells were tolerant to 17-DMAG at bloodstream concentrations as high as 2680 nM [266]. The 17-AAG derivative inhibited mitosis and cytokinesis and sensitized trypanosomes to heat stress [265]. Orally administered 17-DMAG, on the other hand, due to its optimized solubility, cured trypanosome-infected mice models within 5 days [265]. The affinity of geldanamycin and its derivatives, 17-AAG and 17-DMAG, for TbHsp83 was higher than the affinity of these compounds toward the human Hsp90 cytosolic paralogues [263]. The treatment of trypanosomes with 17-AAG phenocopies the knockdown of TbHsp83, having a negative effect on cytokinesis and kinetoplast segregation [249]. Knocking down the mitochondrial T. brucei Hsp90, TbHsp84, results in a loss of the kDNA [249]. Protein phosphatase 5 (PP5) influenced the susceptibility of TbHsp83 and trypanosomes to geldanamycin [264]. PP5, which is also a cytosolic protein, co-localizes with TbHsp83 under stress conditions [264]. The levels of PP5 expression were inversely proportional to the trypanocidal effects of geldanamycin [264]. The knockdown of PP5, particularly in BSFs, is highly detrimental, resulting in an 8-fold decrease in parasitic growth [267].

6.5.8. Adenosine Analogues

T. brucei and other trypanosomatid organisms are incapable of de novo purine biosynthesis and need to salvage them from the host [268,269]. An understanding of the purine salvage machinery enables the development of drug targets [270]. T. brucei has a transport system that imports purines from the host [271]; these transporters include the P2 and P1 adenosine transporter families [272,273]. These transporters could facilitate the delivery of trypanocidal adenosine analogues into the cell [270].
The single P2 transporter, encoded by TbAT1, is responsible for melarsoprol–pentamidine cross resistance and is less desirable to target for adenosine analogue delivery [95,270]. The P1 transporter family, consisting of four members, is redundant in specificity, and trypanocidal adenosine nucleosides that preferentially traverse through them are unlikely to face resistance challenges [270]. Another aspect of these analogues’ attractiveness is that they are likely to penetrate the BBB as it possesses purine transporters [274].
Adenosine analogues such as cordycepin and tubercidin have been identified as trypanocides [275,276]. However, tubercidin is toxic to mammalian cells [276,277]. A hybrid molecule between cordycepin and tubercidin has brought about the 3′-deoxy-7-deazaadenosine analogues, which are effective against trypanosomes in mouse models, with 3-deoxytubercidin being identified as the most promising of the 3′-deoxy-7-deazaadenosine analogue derivatives [269,277]. Derivatives of 3′-deoxy-7-deazaadenosine, referred to as C6-0-alkylated 7-deazainosine analogues, have been identified as trypanocides, with potential to be developed further along the pipeline of trypanosomiasis drugs [278]. Cordycepin, when administered in conjunction with deoxycoformycin or coformycin, has been reported to be curative in mice infected with T. b. brucei, even when the parasites have reached the brain parenchyma [279].

7. Vaccine Development

As a consequence of antigenic variation, successful vaccine development for HAT is deemed unlikely [4]. This is due to the constant switching of surface antigens that could potentially be targets of the vaccine-primed immune response. VSGs also shield potential vaccine antigenic targets such as the antigenic invariant surface glycoproteins (ISGs) from the immune system’s components [280,281,282,283,284,285]. ISGs are transmembrane proteins that intercalate between the VSGs [281]. Additionally, B cell impairment and depletion by trypanosomes means that developing vaccine antigens based on memory B cells could be unlikely [286]. Examples of ISGs are ISG65 and ISG75, which present approximately 70,000 and 50,000 copies of themselves, respectively, on the trypanosome’s surface [183]. Both ISG65 and ISG75 have been studied as potential antigenic targets in T. brucei vaccine development, with ISG75 having been determined to induce insufficient protection [280,281,287]. Interestingly, ISG75 is a target of suramin, implicated in the endocytosis of the drug [282]. In a recent study, the immunogenic potential of ISG75 alongside T. brucei enolase (TbENO) was reassessed. TbENO, like with enolase in T. cruzi and Leishmania, forms part of the secretome and is highly antigenic [288]. Though proving to be highly immunogenic, both ISG75 and TbENO were ultimately determined to be inadequate as antigenic targets.

8. Conclusions

Sleeping sickness is a neglected tropical disease that has remained a global disease burden targeted for elimination as a public health concern by 2030 [289]. However, the interruption in both active and passive screening, especially in countries like the DRC, which currently present the highest number of cases, is likely to affect the predicted elimination time according to prediction models [289,290]. Despite positive strides made by the WHO, non-governmental agencies, and researchers, the COVID-19 pandemic created a major setback. Even with these setbacks, the Ugandan Ministry of Health reported in October 2022 that with individual case management, intense surveillance, and vector control, they had eliminated sleeping sickness caused by T. gambiense as a public health problem [291].
Although the number of reported cases of HAT has drastically reduced over the past few decades, the reports may not be exactly due to poor surveillance in affected areas [33]. Such areas, accounting for “invisible” cases as well as the reemergence of the disease that has occurred in the past [292], call for relentless efforts in the search for updated screening and therapeutics. Improved diagnostic tools are needed to support treatment for tests for a cure in clinical trials and for surveillance of populations in control programs [293]. The most recent drug approved for HAT treatment, fexinidazole, which is administered orally and is efficacious in both stages of the disease, is evidence of sustained efforts to find alternative treatments [129].
This study has summarized most of the promising therapeutics currently being explored and their challenges. The ideal anti-parasitic drug target would be an essential protein for parasite survival significantly different from its orthologue in the host [294]. However, the presence of an orthologue in mammals does not necessarily mean that a particular target should not be pursued [117]. Both CPSF3 and NMT are present in eukaryotic cells and are promising targets for prospective HAT drugs as well as the treatment of cancer [139,170,295,296]. The inhibition of T. brucei N-myristoyltransferase cured trypanosomiasis in mice, showed promising selectivity, and fulfilled most of the requirements for a novel chemotherapeutic agent for HAT [163]. TbHsp90 is similar to its human counterpart but demonstrates biochemical differences that increase inhibitor efficiency.
Another approach involves the use of host-directed strategies, which may include targeting the immune system [297]. In T. brucei, this would be favorable as sustained activation of the immune system may be damaging [117]. In cutaneous leishmaniasis, an immune stimulator referred to as CpG oligonucleotide D35 (CpG ODN D35) has been shown to enhance the response to treatment with the pentavalent antimonial stibogluconate in macaques infected with L. major [298]. The animals presented with relatively smaller lesions that healed relatively quicker when treated with CpG ODN D35 and pentavalent stibogluconate [299]. This serves as an example of how host-targeted approaches could be applied in treating diseases caused by protozoan parasites, particularly kinetoplastids.
It should also be noted that the DNDi has also adopted an artificial intelligence (AI)-driven drug design approach in the quest to discover more therapeutics for neglected diseases. This comes at a time when AI is being employed as one of the tools by which drugs may be designed, developed, and brought to market [299]. Given the status of HAT as an NTD, the potential of AI drug design to lower costs and lessen the time for drug discovery could contribute positively to the development of efficacious HAT treatments [299].
A greater understanding of drug resistance pathways and how they could be circumvented may provide solutions to address resistance to existing drugs. A preemptory research effort into drug resistance mechanisms is needed to identify drug resistance phenotypes as they emerge. Knowledge of persister forms, dormancy, and tissue distribution dynamics is required for the development of more effective chemotherapeutic strategies and to ensure that the drugs can access all sites of infection. With the consistent and concerted efforts of the WHO as well as other public and private partnerships, the eradication of HAT as a public health concern may be a reality. If both COVID-19 and sleeping sickness are viewed as endemic in Africa, then both need to be managed concurrently.

Author Contributions

Conceptualization, A.B.; writing—original draft preparation, M.J. and M.M.; writing—review and editing, A.B. and A.L.E.; visualization, M.M.; supervision, A.B. and A.L.E.; project administration, A.B. and A.L.E. All authors have read and agreed to the published version of the manuscript.

Funding

M.J. is a recipient of the 2018 DAAD In-Region Scholarship, and M.M. is a recipient of the Pearson Young Scholarship. A.L.E. is supported by the South African Research Chairs Initiative (SARChI) of the Department of Science and Innovation (DSI) and National Research Foundation of South Africa (NRF) (Grant No 98566), the NRF Competitive Grants for Rated Researchers (CPRR) (Grant No 129262), and Rhodes University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Raper, J.; Portela, M.P.M.; Lugli, E.; Frevert, U.; Tomlinson, S. Trypanosome lytic factors: Novel mediators of human innate immunity. Curr. Opin. Microbiol. 2001, 4, 402–408. [Google Scholar] [CrossRef] [PubMed]
  2. Lopes, A.H.; Souto-Padrón, T.; Dias, F.A.; Gomes, M.T.; Rodrigues, G.C.; Zimmermann, L.T.; e Silva, T.L.A.; Vermelho, A.B. Trypanosomatids: Odd Organisms, Devastating Diseases. Open Parasitol. J. 2010, 4, 30–59. [Google Scholar] [CrossRef] [Green Version]
  3. Alsan, M. The Effect of the TseTse Fly on African Development. Am. Econ. Rev. 2015, 105, 382–410. [Google Scholar] [CrossRef] [Green Version]
  4. Kennedy, P.G. Clinical features, diagnosis, and treatment of human African trypanosomiasis (sleeping sickness). Lancet Neurol. 2013, 12, 186–194. [Google Scholar] [CrossRef] [PubMed]
  5. Babokhov, P.; Sanyaolu, A.O.; Oyibo, W.A.; Fagbenro-Beyioku, A.F.; Iriemenam, N.C. A current analysis of chemotherapy strategies for the treatment of human African trypanosomiasis. Pathog. Glob. Health 2013, 107, 242–252. [Google Scholar]
  6. Cayla, M.; Rojas, F.; Silvester, E.; Venter, F.; Matthews, K.R. African trypanosomes. Parasit. Vectors 2019, 12, 190. [Google Scholar]
  7. Holmes, P. Tsetse-transmitted trypanosomes—Their biology, disease impact and control. J. Invertebr. Pathol. 2013, 112, S11–S14. [Google Scholar] [CrossRef] [PubMed]
  8. Maudlin, I. African trypanosomiasis. Ann. Trop. Med. Parasitol. 2006, 100, 679–701. [Google Scholar] [CrossRef]
  9. Simarro, P.; Diarra, A.; Ruiz Postigo, J.A.; Franco, J.R.; Jannin, J.G. The Human African Trypanosomiasis Control and Surveillance Programme of the World Health Organization 2000–2009: The Way Forward. PLoS Negl. Trop. Dis. 2011, 5, e1007. [Google Scholar]
  10. Gruvel, J. Considérations générales sur la signification de la transmission mécanique des trypanosomoses chez le bétail. Int. J. Trop. Insect Sci. 1980, 1, 55–57. [Google Scholar] [CrossRef]
  11. Lindner, A.K.; Priotto, G. The Unknown Risk of Vertical Transmission in Sleeping Sickness—A Literature Review. PLoS Negl. Trop. Dis. 2010, 4, e783. [Google Scholar] [CrossRef] [PubMed]
  12. Simarro, P.; Franco, J.; Diarra, A.; Jannin, J. Epidemiology of human African trypanosomiasis. Clin. Epidemiol. 2014, 6, 257–275. [Google Scholar] [CrossRef] [PubMed]
  13. Kennedy, P.G.E. Human African trypanosomiasis of the CNS: Current issues and challenges. J. Clin. Investig. 2004, 113, 496–504. [Google Scholar]
  14. Kennedy, P.G.E.; Rodgers, J. Clinical and Neuropathogenetic Aspects of Human African Trypanosomiasis. Front. Immunol. 2019, 10, 39. [Google Scholar] [CrossRef] [Green Version]
  15. Torr, S.J.; Chamisa, A.; Mangwiro, T.N.C.; Vale, G.A. Where, When and Why Do Tsetse Contact Humans? Answers from Studies in a National Park of Zimbabwe. PLoS Negl. Trop. Dis. 2012, 6, e1791. [Google Scholar] [CrossRef] [Green Version]
  16. Simo, G.; Mbida, J.; Eyenga, V.; Asonganyi, T.; Njiokou, F.; Grébaut, P. Challenges towards the elimination of Human African Trypanosomiasis in the sleeping sickness focus of Campo in southern Cameroon. Parasit. Vectors 2014, 7, 374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Berrang-Ford, L.; Odiit, M.; Maiso, F.; Waltner-Toews, D.; McDermott, J. Sleeping sickness in Uganda: Revisiting current and historical distributions. Afr. Health Sci. 2006, 6, 223–231. [Google Scholar]
  18. Berrang-Ford, L.; Wamboga, C.; Kakembo, A.S.L. Trypanososma brucei rhodesiense Sleeping Sickness, Uganda. Emerg. Infect. Dis. 2012, 18, 1686–1687. [Google Scholar] [CrossRef]
  19. Wamwiri, F.N.; Changasi, R.E. Tsetse Flies (Glossina) as Vectors of Human African Trypanosomiasis: A Review. BioMed Res. Int. 2016, 2016, 6201350. [Google Scholar] [CrossRef] [Green Version]
  20. Tirados, I.; Esterhuizen, J.; Kovacic, V.; Mangwiro, T.N.C.; Vale, G.A.; Hastings, I.; Solano, P.; Lehane, M.J.; Torr, S.J. Tsetse Control and Gambian Sleeping Sickness; Implications for Control Strategy. PLoS Negl. Trop. Dis. 2015, 9, e0003822. [Google Scholar] [CrossRef] [Green Version]
  21. Egeru, A.; Opio, J.; Siya, A.; Barasa, B.; Magaya, J.P.; Namaalwa, J.J. Tsetse Invasion as an Emerging Threat to Socioecological Resilience of Pastoral Communities in Karamoja, Uganda. Sustainability 2020, 12, 1599. [Google Scholar] [CrossRef] [Green Version]
  22. Louis, F.J.; Simarro, P.P. Rough start for the fight against sleeping sickness in French equatorial Africa. Med. Trop. Rev. Corps Sante Colon. 2005, 65, 251–257. [Google Scholar]
  23. Steverding, D. The history of African trypanosomiasis. Parasit. Vectors 2008, 1, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Hide, G. History of Sleeping Sickness in East Africa. Clin. Microbiol. Rev. 1999, 12, 112–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Brun, R.; Blum, J.; Chappuis, F.; Burri, C. Human African trypanosomiasis. Lancet 2010, 375, 148–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Smith, D.H.; Pepin, J.; Stich, A.H.R. Human African trypanosomiasis: An emerging public health crisis. Br. Med. Bull. 1998, 54, 341–355. [Google Scholar] [CrossRef] [Green Version]
  27. Moore, A.; Richer, M. Re-emergence of epidemic sleeping sickness in southern Sudan. Trop. Med. Int. Health 2001, 6, 342–347. [Google Scholar] [CrossRef] [Green Version]
  28. Nieuwenhove, S.V.; Betu-Ku-Mesu, V.K.; Diabakana, P.M.; Declercq, J.; Bilenge, C.M.M. Sleeping sickness resurgence in the DRC: The past decade. Trop. Med. Int. Health 2001, 6, 335–341. [Google Scholar] [CrossRef]
  29. Stanghellini, A.; Josenando, T. The situation of sleeping sickness in Angola: A calamity. Trop. Med. Int. Health 2001, 6, 330–334. [Google Scholar] [CrossRef] [Green Version]
  30. Franco, J.R.; Simarro, P.P.; Diarra, A.; Ruiz-Postigo, J.A.; Jannin, J.G. The journey towards elimination of gambiense human African trypanosomiasis: Not far, nor easy. Parasitology 2014, 141, 748–760. [Google Scholar] [CrossRef]
  31. WHO. Trypanosomiasis, Human African (Sleeping Sickness). 2022. Available online: https://www.who.int/news-room/fact-sheets/detail/trypanosomiasis-human-african-(sleeping-sickness) (accessed on 21 May 2022).
  32. Simarro, P.; Cecchi, G.; Franco, J.; Paone, M.; Diarra, A.; Ruiz-Postigo, J.; Fèvre, E.; Mattioli, R.; Jannin, M.; Ndung’u, J. Estimating and Mapping the Population at Risk of Sleeping Sickness. PLoS Negl. Trop. Dis. 2012, 6, e1859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Chappuis, F.; Lima, M.A.; Flevaud, L.; Ritmeijer, K. Human African Trypanosomiasis in Areas without Surveillance. Emerg. Infect. Dis. 2010, 16, 354–356. [Google Scholar] [CrossRef] [PubMed]
  34. Ripamonti, D.; Massari, M.; Claudio, A.; Ermanno, G.; Claudio, F.; Brini, M.; Capatti, C.; Suter, F. African Sleeping Sickness in Tourists Returning from Tanzania: The First 2 Italian Cases from a Small Outbreak among European Travelers. Clin. Infect. Dis. 2002, 34, e18–e22. [Google Scholar] [PubMed] [Green Version]
  35. Meehl, G.; Stocker, T.; Collins, W.; Friedlingstein, P.; Gaye, T.; Gregory, J.; Kitoh, A.; Knutti, R.; Murphy, J.; Noda, A.; et al. IPCC, 2007: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK, 2007; pp. 747–846. [Google Scholar]
  36. Urbaniak, M.D.; Guther, M.L.S.; Ferguson, M.A.J. Comparative SILAC proteomic analysis of Trypanosoma brucei bloodstream and procyclic lifecycle stages. PLoS ONE 2012, 7, e36619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Cox, F.E.G. History of sleeping sickness (African trypanosomiasis). Infect. Dis. Clin. N. Am. 2004, 18, 231–245. [Google Scholar] [CrossRef]
  38. Büscher, P.; Cecchi, G.; Jamonneau, V.; Priotto, G. Human African trypanosomiasis. Lancet 2017, 390, 2397–2409. [Google Scholar] [CrossRef]
  39. Chappuis, F.; Loutan, L.; Simarro, P.; Lejon, V.; Büscher, P. Options for Field Diagnosis of Human African Trypanosomiasis. Clin. Microbiol. Rev. 2005, 18, 133–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Cappello, M.; Li, S.; Chen, X.; Li, C.-B.; Harrison, L.; Narashimhan, S.; Beard, C.B.; Aksoy, S. Tsetse thrombin inhibitor: Bloodmeal-induced expression of an anticoagulant in salivary glands and gut tissue of Glossina morsitans morsitans. Proc. Natl. Acad. Sci. USA 1998, 95, 14290–14295. [Google Scholar] [CrossRef]
  41. MacGregor, P.; Savill, N.J.; Hall, D.; Matthews, K.R. Transmission Stages Dominate Trypanosome Within-Host Dynamics during Chronic Infections. Cell Host Microbe 2011, 9, 310–318. [Google Scholar] [CrossRef] [Green Version]
  42. Reuner, B.; Vassella, E.; Yutzy, B.; Boshart, M. Cell density triggers slender to stumpy differentiation of Trypanosoma brucei bloodstream forms in culture. Mol. Biochem. Parasitol. 1997, 90, 269–280. [Google Scholar] [CrossRef]
  43. Seed, J.; Wenck, M. Role of the long slender to short stumpy transition in the life cycle of the african trypanosomes. Kinetoplastid Biol. Dis. 2003, 2, 3. [Google Scholar] [CrossRef] [Green Version]
  44. Mant, M.J.; Parker, K.R. Two Platelet Aggregation Inhibitors in Tsetse (Glossina) Saliva with Studies of Roles of Thrombin and Citrate in in Vitro Platelet Aggregation. Br. J. Haematol. 2008, 48, 601–608. [Google Scholar] [CrossRef] [PubMed]
  45. Caljon, G.; De Ridder, K.; De Baetselier, P.; Coosemans, M.; Van Den Abbeele, J. Identification of a Tsetse Fly Salivary Protein with Dual Inhibitory Action on Human Platelet Aggregation. PLoS ONE 2010, 5, e9671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Zhao, X.; Silva, T.L.A.E.; Cronin, L.; Savage, A.F.; O’Neill, M.; Nerima, B.; Okedi, L.M.; Aksoy, S. Immunogenicity and Serological Cross-Reactivity of Saliva Proteins among Different Tsetse Species. PLoS Negl. Trop. Dis. 2015, 9, e0004038. [Google Scholar] [CrossRef]
  47. Trindade, S.; Rijo-Ferreira, F.; Carvalho, T.; Pinto-Neves, D.; Guegan, F.; Aresta-Branco, F.; Bento, F.; Young, S.A.; Pinto, A.; Van Den Abbeele, J.; et al. Trypanosoma brucei Parasites Occupy and Functionally Adapt to the Adipose Tissue in Mice. Cell Host Microbe 2016, 19, 837–848. [Google Scholar] [CrossRef] [Green Version]
  48. Hoare, C.A.; Wallace, F.G. Developmental Stages of Trypanosomatid Flagellates: A New Terminology. Nature 1966, 212, 1385–1386. [Google Scholar] [CrossRef]
  49. Wheeler, R.J.; Gluenz, E.; Gull, K. The Limits on Trypanosomatid Morphological Diversity. PLoS ONE 2013, 8, e79581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Losos, G.J.; Ikede, B.O. Review of Pathology of Diseases in Domestic and Laboratory Animals Caused by Trypanosoma congolense, T. vivax, T. brucei, T. rhodesiense and T. gambiense. Vet. Pathol. 1972, 9, 1–79. [Google Scholar] [CrossRef] [Green Version]
  51. Caljon, G.; Van Reet, N.; De Trez, C.; Vermeersch, M.; Pérez-Morga, D.; Van Den Abbeele, J. The Dermis as a Delivery Site of Trypanosoma brucei for Tsetse Flies. PLoS Pathog. 2016, 12, e1005744. [Google Scholar] [CrossRef] [Green Version]
  52. Capewell, P.; Cren-Travaillé, C.; Marchesi, F.; Johnston, P.; Clucas, C.; Benson, R.A.; Gorman, T.-A.; Calvo-Alvarez, E.; Crouzols, A.; Jouvion, G.; et al. The skin is a significant but overlooked anatomical reservoir for vector-borne African trypanosomes. eLife 2016, 5, e17716. [Google Scholar] [CrossRef]
  53. Tanowitz, H.B.; Scherer, P.E.; Mota, M.M.; Figueiredo, L.M. Adipose Tissue: A Safe Haven for Parasites? Trends Parasitol. 2017, 33, 276–284. [Google Scholar] [PubMed] [Green Version]
  54. Poudyal, N.R.; Paul, K.S. Fatty acid uptake in Trypanosoma brucei: Host resources and possible mechanisms. Front. Cell. Infect. Microbiol. 2022, 12, 949409. [Google Scholar] [PubMed]
  55. Claes, F.; Vodnala, S.K.; van Reet, N.; Boucher, N.; Lunden-Miguel, H.; Baltz, T.; Goddeeris, B.M.; Büscher, P.; Rottenberg, M.E. Bioluminescent Imaging of Trypanosoma brucei Shows Preferential Testis Dissemination Which May Hamper Drug Efficacy in Sleeping Sickness. PLoS Negl. Trop. Dis. 2009, 3, e486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Biteau, N.; Asencio, C.; Izotte, J.; Rousseau, B.; Fèvre, M.; Pillay, D.; Baltz, T. Trypanosoma brucei gambiense Infections in Mice Lead to Tropism to the Reproductive Organs, and Horizontal and Vertical Transmission. PLoS Negl. Trop. Dis. 2016, 10, e0004350. [Google Scholar] [CrossRef]
  57. Butter, F.; Bucerius, F.; Michel, M.; Cicova, Z.; Mann, M.; Janzen, C.J. Comparative proteomics of two life cycle stages of stable isotope-labeled Trypanosoma brucei reveals novel components of the parasite’s host adaptation machinery. Mol. Cell. Proteomics MCP 2013, 12, 172–179. [Google Scholar] [CrossRef] [Green Version]
  58. Szöör, B.; Haanstra, J.R.; Gualdrón-López, M.; Michels, P.A. Evolution, dynamics and specialized functions of glycosomes in metabolism and development of trypanosomatids. Curr. Opin. Microbiol. 2014, 22, 79–87. [Google Scholar] [CrossRef] [Green Version]
  59. Stuart, K.; Brun, R.; Croft, S.; Fairlamb, A.; Gürtler, R.E.; McKerrow, J.; Reed, S.; Tarleton, R. Kinetoplastids: Related protozoan pathogens, different diseases. J. Clin. Investig. 2008, 118, 1301–1310. [Google Scholar] [CrossRef] [Green Version]
  60. MacLean, L.M.; Odiit, M.; Chisi, J.E.; Kennedy, P.G.E.; Sternberg, J.M. Focus–Specific Clinical Profiles in Human African Trypanosomiasis Caused by Trypanosoma brucei rhodesiense. PLoS Negl. Trop. Dis. 2010, 4, e906. [Google Scholar] [CrossRef] [Green Version]
  61. Kennedy, P.G.E. The continuing problem of human African trypanosomiasis (sleeping sickness). Ann. Neurol. 2008, 64, 116–126. [Google Scholar] [CrossRef]
  62. Bisser, S.; Lejon, V.; Preux, P.M.; Bouteille, B.; Stanghellini, A.; Jauberteau, M.O.; Büscher, P.; Dumas, M. Blood–cerebrospinal fluid barrier and intrathecal immunoglobulins compared to field diagnosis of central nervous system involvement in sleeping sickness. J. Neurol. Sci. 2002, 193, 127–135. [Google Scholar] [CrossRef]
  63. Buguet, A.; Bisser, S.; Josenando, T.; Chapotot, F.; Cespuglio, R. Sleep structure: A new diagnostic tool for stage determination in sleeping sickness. Acta Trop. 2005, 93, 107–117. [Google Scholar] [CrossRef] [PubMed]
  64. Bonnet, J.; Boudot, C.; Courtioux, B. Overview of the Diagnostic Methods Used in the Field for Human African Trypanosomiasis: What Could Change in the Next Years? BioMed Res. Int. 2015, 2015, 583262. [Google Scholar] [PubMed] [Green Version]
  65. Stich, A. Human African trypanosomiasis. BMJ 2002, 325, 203–206. [Google Scholar] [CrossRef] [PubMed]
  66. Ponte-Sucre, A. An Overview of Trypanosoma brucei Infections: An Intense Host–Parasite Interaction. Front. Microbiol. 2016, 7, 2126. [Google Scholar] [PubMed] [Green Version]
  67. de Raadt, P.; Koten, J.W. Myocarditis in Rhodesiense trypanosomiasis. East Afr. Med. J. 1968, 45, 128–132. [Google Scholar]
  68. Petzke, F.; Heppner, C.; Mbulamberi, D.; Winkelmann, W.; Chrousos, G.P.; Allolio, B.; Reincke, M. Hypogonadism in Rhodesian sleeping sickness: Evidence for acute and chronic dysfunction of the hypothalamic-pituitary-gonadal axis. Fertil. Steril. 1996, 65, 68–75. [Google Scholar]
  69. Kuepfer, I.; Hhary, E.; Allan, M.; Edielu, A.; Burri, C.; Blum, J. Clinical Presentation of T.b. rhodesiense Sleeping Sickness in Second Stage Patients from Tanzania and Uganda. PLoS Negl. Trop. Dis. 2011, 5, e968. [Google Scholar] [CrossRef] [Green Version]
  70. Büscher, P.; Mumba Ngoyi, D.; Kaboré, J.; Lejon, V.; Robays, J.; Jamonneau, V.; Bebronne, N.; Van der Veken, W.; Biéler, S. Improved Models of Mini Anion Exchange Centrifugation Technique (mAECT) and Modified Single Centrifugation (MSC) for Sleeping Sickness Diagnosis and Staging. PLoS Negl. Trop. Dis. 2009, 3, e471. [Google Scholar] [CrossRef] [Green Version]
  71. Camara, M.; Camara, O.; Ilboudo, H.; Sakande, H.; Kaboré, J.; N’Dri, L.; Jamonneau, V.; Bucheton, B. Sleeping sickness diagnosis: Use of buffy coats improves the sensitivity of the mini anion exchange centrifugation test: mAECT-bc and sleeping sickness diagnostic. Trop. Med. Int. Health 2010, 15, 796–799. [Google Scholar]
  72. Truc, P.; Lejon, V.; Magnus, E.; Jamonneau, V.; Nangouma, A.; Verloo, D.; Penchenier, L.; Büscher, P. Evaluation of the micro-CATT, CATT/Trypanosoma brucei gambiense, and LATEX/T b gambiense methods for serodiagnosis and surveillance of human African trypanosomiasis in West and Central Africa. Bull. World Health Organ. 2002, 80, 882–886. [Google Scholar]
  73. Bisser, S.; Lumbala, C.; Nguertoum, E.; Kande, V.; Flevaud, L.; Vatunga, G.; Boelaert, M.; Büscher, P.; Josenando, T.; Bessell, P.R.; et al. Sensitivity and Specificity of a Prototype Rapid Diagnostic Test for the Detection of Trypanosoma brucei gambiense Infection: A Multi-centric Prospective Study. PLoS Negl. Trop. Dis. 2016, 10, e0004608. [Google Scholar]
  74. Bisser, S.; N’Siesi, F.; Lejon, V.; Preux, P.; Van Nieuwenhove, S.; Miaka Mia Bilenge, C.; Büscher, P. Equivalence Trial of Melarsoprol and Nifurtimox Monotherapy and Combination Therapy for the Treatment of Second-Stage Trypanosoma brucei gambiense Sleeping Sickness. J. Infect. Dis. 2007, 195, 322–329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Patterson, S.; Wyllie, S. Nitro drugs for the treatment of trypanosomatid diseases: Past, present, and future prospects. Trends Parasitol. 2014, 30, 289–298. [Google Scholar]
  76. Kansiime, F.; Adibaku, S.; Wamboga, C.; Idi, F.; Kato, C.D.; Yamuah, L.; Vaillant, M.; Kioy, D.; Olliaro, P.; Matovu, E. A multicentre, randomised, non-inferiority clinical trial comparing a nifurtimox-eflornithine combination to standard eflornithine monotherapy for late stage Trypanosoma brucei gambiense human African trypanosomiasis in Uganda. Parasit. Vectors 2018, 11, 105. [Google Scholar]
  77. Munoz-Calderon, A.; Díaz-Bello, Z.; Ramírez, J.; Noya, O.; de Noya, B. Nifurtimox response of Trypanosoma cruzi isolates from an outbreak of Chagas disease in Caracas, Venezuela. J. Vector Borne Dis. 2019, 56, 237. [Google Scholar]
  78. Masocha, W.; Rottenberg, M.E.; Kristensson, K. Migration of African trypanosomes across the blood–brain barrier. Physiol. Behav. 2007, 92, 110–114. [Google Scholar] [CrossRef]
  79. Priotto, G.; Pinoges, L.; Fursa, I.B.; Burke, B.; Nicolay, N.; Grillet, G.; Hewison, C.; Balasegaram, M. Safety and effectiveness of first line eflornithine for Trypanosoma brucei gambiense sleeping sickness in Sudan: Cohort study. BMJ 2008, 336, 705–708. [Google Scholar] [PubMed] [Green Version]
  80. Sanderson, L.; Dogruel, M.; Rodgers, J.; De Koning, H.P.; Thomas, S.A. Pentamidine Movement across the Murine Blood-Brain and Blood-Cerebrospinal Fluid Barriers: Effect of Trypanosome Infection, Combination Therapy, P-Glycoprotein, and Multidrug Resistance-Associated Protein. J. Pharmacol. Exp. Ther. 2009, 329, 967–977. [Google Scholar] [PubMed]
  81. Singaro, J.R.S.; Yapo, F.B.; Doua, F.; Miezan, T.W.; Baltz, T. The Efficacy of Pentamidine in the Treatment of Early-Late Stage Trypanosoma brucei gambiense Trypanosomiasis *. Am. J. Trop. Med. Hyg. 1996, 55, 586–588. [Google Scholar]
  82. Sekhar, G.N.; Georgian, A.R.; Sanderson, L.; Vizcay-Barrena, G.; Brown, R.C.; Muresan, P.; Fleck, R.A.; Thomas, S.A. Organic cation transporter 1 (OCT1) is involved in pentamidine transport at the human and mouse blood-brain barrier (BBB). PLoS ONE 2017, 12, e0173474. [Google Scholar] [CrossRef] [Green Version]
  83. Gehrig, S.; Efferth, T. Development of drug resistance in Trypanosoma brucei rhodesiense and Trypanosoma brucei gambiense. Treatment of human African trypanosomiasis with natural products (Review). Int. J. Mol. Med. 2008, 22, 411–419. [Google Scholar]
  84. Barrett, M.P.; Boykin, D.W.; Brun, R.; Tidwell, R.R. Human African trypanosomiasis: Pharmacological re-engagement with a neglected disease: Drugs for human African trypanosomiasis. Br. J. Pharmacol. 2007, 152, 1155–1171. [Google Scholar]
  85. Van Bogaert, I.; Haemers, A. Eflornithine: A new drug in the treatment of sleeping sickness. Pharm. Weekbl. 1989, 11, 69–75. [Google Scholar]
  86. Barrett, S.V.; Barrett, M.P. Anti-sleeping sickness drugs and cancer chemotherapy. Parasitol. Today Pers. Ed 2000, 16, 7–9. [Google Scholar] [CrossRef] [PubMed]
  87. Fonseca, M.S.; Comini, M.A.; Resende, B.V.; Santi, A.M.M.; Zoboli, A.P.; Moreira, D.S.; Murta, S.M.F. Ornithine decarboxylase or gamma-glutamylcysteine synthetase overexpression protects Leishmania (Vianna) guyanensis against antimony. Exp. Parasitol. 2017, 175, 36–43. [Google Scholar] [PubMed]
  88. Ariyanayagam, M.R.; Oza, S.L.; Guther, M.L.S.; Fairlamb, A.H. Phenotypic analysis of trypanothione synthetase knockdown in the African trypanosome. Biochem. J. 2005, 391, 425–432. [Google Scholar] [CrossRef]
  89. Meyskens, F.L.; Gerner, E.W. Development of difluoromethylornithine (DFMO) as a chemoprevention agent. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 1999, 5, 945–951. [Google Scholar]
  90. Priotto, G.; Kasparian, S.; Mutombo, W.; Ngouama, D.; Ghorashian, S.; Arnold, U.; Ghabri, S.; Baudin, E.; Buard, V.; Kazadi-Kyanza, S.; et al. Nifurtimox-eflornithine combination therapy for second-stage African Trypanosoma brucei gambiense trypanosomiasis: A multicentre, randomised, phase III, non-inferiority trial. Lancet 2009, 374, 56–64. [Google Scholar]
  91. Schmidt, R.S.; Macêdo, J.P.; Steinmann, M.E.; Salgado, A.G.; Bütikofer, P.; Sigel, E.; Rentsch, D.; Mäser, P. Transporters of Trypanosoma brucei —Phylogeny, physiology, pharmacology. FEBS J. 2018, 285, 1012–1023. [Google Scholar] [CrossRef] [Green Version]
  92. Hassan, H.F.; Coombs, G.H. Purine and pyrimidine metabolism in parasitic protozoa. FEMS Microbiol. Rev. 1988, 4, 47–83. [Google Scholar]
  93. Chaudhary, K.; Roos, D.S. Protozoan genomics for drug discovery. Nat. Biotechnol. 2005, 23, 1089–1091. [Google Scholar] [CrossRef]
  94. Alsford, S.; Field, M.C.; Horn, D. Receptor-mediated endocytosis for drug delivery in African trypanosomes: Fulfilling Paul Ehrlich’s vision of chemotherapy. Trends Parasitol. 2013, 29, 207–212. [Google Scholar]
  95. Bridges, D.J.; Gould, M.K.; Nerima, B.; Mäser, P.; Burchmore, R.J.S.; De Koning, H.P. Loss of the High-Affinity Pentamidine Transporter Is Responsible for High Levels of Cross-Resistance between Arsenical and Diamidine Drugs in African Trypanosomes. Mol. Pharmacol. 2007, 71, 1098–1108. [Google Scholar] [CrossRef] [Green Version]
  96. Baker, N.; Glover, L.; Munday, J.C.; Aguinaga Andrés, D.; Barrett, M.P.; de Koning, H.P.; Horn, D. Aquaglyceroporin 2 controls susceptibility to melarsoprol and pentamidine in African trypanosomes. Proc. Natl. Acad. Sci. USA 2012, 109, 10996–11001. [Google Scholar] [CrossRef]
  97. Matovu, E.; Stewart, M.L.; Geiser, F.; Brun, R.; Mäser, P.; Wallace, L.J.M.; Burchmore, R.J.; Enyaru, J.C.K.; Barrett, M.P.; Kaminsky, R.; et al. Mechanisms of arsenical and diamidine uptake and resistance in Trypanosoma brucei. Eukaryot. Cell 2003, 2, 1003–1008. [Google Scholar] [CrossRef] [Green Version]
  98. Munday, J.C.; Eze, A.A.; Baker, N.; Glover, L.; Clucas, C.; Aguinaga Andres, D.; Natto, M.J.; Teka, I.A.; McDonald, J.; Lee, R.S.; et al. Trypanosoma brucei aquaglyceroporin 2 is a high-affinity transporter for pentamidine and melaminophenyl arsenic drugs and the main genetic determinant of resistance to these drugs. J. Antimicrob. Chemother. 2014, 69, 651–663. [Google Scholar] [CrossRef] [PubMed]
  99. De Koning, H.P. Uptake of Pentamidine in Trypanosoma brucei brucei is Mediated by Three Distinct Transporters: Implications for Cross-Resistance with Arsenicals. Mol. Pharmacol. 2001, 59, 586–592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Munday, J.C.; Settimo, L.; De Koning, H.P. Transport proteins determine drug sensitivity and resistance in a protozoan parasite, Trypanosoma brucei. Front. Pharmacol. 2015, 6, 32. [Google Scholar] [CrossRef] [PubMed]
  101. Bassarak, B.; Uzcátegui, N.L.; Schönfeld, C.; Duszenko, M. Functional Characterization of Three Aquaglyceroporins from Trypanosoma brucei in Osmoregulation and Glycerol Transport. Cell. Physiol. Biochem. 2011, 27, 411–420. [Google Scholar] [CrossRef]
  102. Pyana Pati, P.; Van Reet, N.; Mumba Ngoyi, D.; Ngay Lukusa, I.; Karhemere Bin Shamamba, S.; Büscher, P. Melarsoprol Sensitivity Profile of Trypanosoma brucei gambiense Isolates from Cured and Relapsed Sleeping Sickness Patients from the Democratic Republic of the Congo. PLoS Negl. Trop. Dis. 2014, 8, e3212. [Google Scholar] [CrossRef] [Green Version]
  103. Graf, F.E.; Baker, N.; Munday, J.C.; De Koning, H.P.; Horn, D.; Mäser, P. Chimerization at the AQP2–AQP3 locus is the genetic basis of melarsoprol–pentamidine cross-resistance in clinical Trypanosoma brucei gambiense isolates. Int. J. Parasitol. Drugs Drug Resist. 2015, 5, 65–68. [Google Scholar] [CrossRef] [Green Version]
  104. Quintana, J.F.; Bueren-Calabuig, J.; Zuccotto, F.; De Koning, H.P.; Horn, D.; Field, M.C. Instability of aquaglyceroporin (AQP) 2 contributes to drug resistance in Trypanosoma brucei. PLoS Negl. Trop. Dis. 2020, 14, e0008458. [Google Scholar] [CrossRef] [PubMed]
  105. Alghamdi, A.H.; Munday, J.C.; Campagnaro, G.D.; Gurvic, D.; Svensson, F.; Okpara, C.E.; Kumar, A.; Quintana, J.; Martin Abril, M.E.; Milić, P.; et al. Positively selected modifications in the pore of TbAQP2 allow pentamidine to enter Trypanosoma brucei. eLife 2020, 9, e56416. [Google Scholar] [CrossRef] [PubMed]
  106. Song, J.; Baker, N.; Rothert, M.; Henke, B.; Jeacock, L.; Horn, D.; Beitz, E. Pentamidine Is Not a Permeant but a Nanomolar Inhibitor of the Trypanosoma brucei Aquaglyceroporin-2. PLoS Pathog. 2016, 12, e1005436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Unciti-Broceta, J.D.; Arias, J.L.; Maceira, J.; Soriano, M.; Ortiz-González, M.; Hernández-Quero, J.; Muñóz-Torres, M.; De Koning, H.P.; Magez, S.; Garcia-Salcedo, J.A. Specific Cell Targeting Therapy Bypasses Drug Resistance Mechanisms in African Trypanosomiasis. PLoS Pathog. 2015, 11, e1004942. [Google Scholar] [CrossRef]
  108. Stijlemans, B.; Conrath, K.; Cortez-Retamozo, V.; Van Xong, H.; Wyns, L.; Senter, P.; Revets, H.; De Baetselier, P.; Muyldermans, S.; Magez, S. Efficient Targeting of Conserved Cryptic Epitopes of Infectious Agents by Single Domain Antibodies. J. Biol. Chem. 2004, 279, 1256–1261. [Google Scholar] [CrossRef] [Green Version]
  109. Jeacock, L.; Baker, N.; Wiedemar, N.; Mäser, P.; Horn, D. Aquaglyceroporin-null trypanosomes display glycerol transport defects and respiratory-inhibitor sensitivity. PLoS Pathog. 2017, 13, e1006307. [Google Scholar] [CrossRef] [Green Version]
  110. Zoltner, M.; Campagnaro, G.D.; Taleva, G.; Burrell, A.; Cerone, M.; Leung, K.-F.; Achcar, F.; Horn, D.; Vaughan, S.; Gadelha, C.; et al. Suramin exposure alters cellular metabolism and mitochondrial energy production in African trypanosomes. J. Biol. Chem. 2020, 295, 8331–8347. [Google Scholar] [CrossRef]
  111. Wiedemar, N.; Graf, F.E.; Zwyer, M.; Ndomba, E.; Kunz Renggli, C.; Cal, M.; Schmidt, R.S.; Wenzler, T.; Mäser, P. Beyond immune escape: A variant surface glycoprotein causes suramin resistance in Trypanosoma brucei: Suramin resistance in T. brucei. Mol. Microbiol. 2018, 107, 57–67. [Google Scholar] [CrossRef] [Green Version]
  112. Zeelen, J.; Van Straaten, M.; Verdi, J.; Hempelmann, A.; Hashemi, H.; Perez, K.; Jeffrey, P.D.; Hälg, S.; Wiedemar, N.; Mäser, P.; et al. Structure of trypanosome coat protein VSGsur and function in suramin resistance. Nat. Microbiol. 2021, 6, 392–400. [Google Scholar] [CrossRef] [PubMed]
  113. Wiedemar, N.; Zwyer, M.; Zoltner, M.; Cal, M.; Field, M.C.; Mäser, P. Expression of a specific variant surface glycoprotein has a major impact on suramin sensitivity and endocytosis in Trypanosoma brucei. FASEB BioAdv. 2019, 1, 595–608. [Google Scholar] [CrossRef] [Green Version]
  114. Mutuku, C.N.; Bateta, R.; Rono, M.K.; Njunge, J.M.; Awuoche, E.O.; Ndung’u, K.; Mang’era, C.M.; Akoth, M.O.; Adung’a, V.O.; Ondigo, B.N.; et al. Physiological and proteomic profiles of Trypanosoma brucei rhodesiense parasite isolated from suramin responsive and non-responsive HAT patients in Busoga, Uganda. Int. J. Parasitol. Drugs Drug Resist. 2021, 15, 57–67. [Google Scholar] [CrossRef]
  115. Vincent, I.M.; Creek, D.; Watson, D.G.; Kamleh, M.A.; Woods, D.J.; Wong, P.E.; Burchmore, R.J.S.; Barrett, M.P. A molecular mechanism for eflornithine resistance in African trypanosomes. PLoS Pathog. 2010, 6, e1001204. [Google Scholar] [CrossRef] [Green Version]
  116. Crilly, N.P.; Mugnier, M.R. Thinking outside the blood: Perspectives on tissue-resident Trypanosoma brucei. PLoS Pathog. 2021, 17, e1009866. [Google Scholar] [CrossRef]
  117. De Rycker, M.; Wyllie, S.; Horn, D.; Read, K.D.; Gilbert, I.H. Anti-trypanosomatid drug discovery: Progress and challenges. Nat. Rev. Microbiol. 2023, 21, 35–50. [Google Scholar] [CrossRef]
  118. Barrett, M.P.; Kyle, D.E.; Sibley, L.D.; Radke, J.B.; Tarleton, R.L. Protozoan persister-like cells and drug treatment failure. Nat. Rev. Microbiol. 2019, 17, 607–620. [Google Scholar] [CrossRef] [PubMed]
  119. Wood, T.K.; Knabel, S.J.; Kwan, B.W. Bacterial persister cell formation and dormancy. Appl. Environ. Microbiol. 2013, 79, 7116–7121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Cerutti, A.; Blanchard, N.; Besteiro, S. The Bradyzoite: A Key Developmental Stage for the Persistence and Pathogenesis of Toxoplasmosis. Pathogens 2020, 9, 234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Francisco, A.F.; Jayawardhana, S.; Lewis, M.D.; Taylor, M.C.; Kelly, J.M. Biological factors that impinge on Chagas disease drug development. Parasitology 2017, 144, 1871–1880. [Google Scholar] [CrossRef]
  122. Sánchez-Valdéz, F.J.; Padilla, A.; Wang, W.; Orr, D.; Tarleton, R.L. Spontaneous dormancy protects Trypanosoma cruzi during extended drug exposure. eLife 2018, 7, e34039. [Google Scholar] [CrossRef]
  123. Silva Pereira, S.; Trindade, S.; De Niz, M.; Figueiredo, L.M. Tissue tropism in parasitic diseases. Open Biol. 2019, 9, 190036. [Google Scholar] [CrossRef] [Green Version]
  124. Trindade, S.; De Niz, M.; Costa-Sequeira, M.; Bizarra-Rebelo, T.; Bento, F.; Dejung, M.; Narciso, M.V.; López-Escobar, L.; Ferreira, J.; Butter, F.; et al. Slow growing behavior in African trypanosomes during adipose tissue colonization. Nat. Commun. 2022, 13, 7548. [Google Scholar] [CrossRef]
  125. Machado, H.; Bizarra-Rebelo, T.; Costa-Sequeira, M.; Trindade, S.; Carvalho, T.; Rijo-Ferreira, F.; Rentroia-Pacheco, B.; Serre, K.; Figueiredo, L.M. Trypanosoma brucei triggers a broad immune response in the adipose tissue. PLoS Pathog. 2021, 17, e1009933. [Google Scholar] [CrossRef]
  126. Reuter, C.; Imdahl, F.; Hauf, L.; Vafadarnejad, E.; Fey, P.; Finger, T.; Walles, H.; Saliba, A.-E.; Groeber-Becker, F.; Engstler, M. Vector-borneTrypanosoma brucei parasites develop in artificial human skin and persist as skin tissue forms. bioRxiv 2021. [Google Scholar] [CrossRef]
  127. Deeks, E.D. Fexinidazole: First Global Approval. Drugs 2019, 79, 215–220. [Google Scholar] [CrossRef] [PubMed]
  128. Hidalgo, J.; Ortiz, J.F.; Fabara, S.P.; Eissa-Garcés, A.; Reddy, D.; Collins, K.D.; Tirupathi, R. Efficacy and Toxicity of Fexinidazole and Nifurtimox Plus Eflornithine in the Treatment of African Trypanosomiasis: A Systematic Review. Cureus 2021, 13, e16881. [Google Scholar] [PubMed]
  129. Kande Betu Ku Mesu, V.; Mutombo Kalonji, W.; Bardonneau, C.; Valverde Mordt, O.; Ngolo Tete, D.; Blesson, S.; Simon, F.; Delhomme, S.; Bernhard, S.; Mahenzi Mbembo, H.; et al. Oral fexinidazole for stage 1 or early stage 2 African Trypanosoma brucei gambiense trypanosomiasis: A prospective, multicentre, open-label, cohort study. Lancet Glob. Health 2021, 9, e999–e1008. [Google Scholar]
  130. Wilkinson, S.R.; Taylor, M.C.; Horn, D.; Kelly, J.M.; Cheeseman, I. A mechanism for cross-resistance to nifurtimox and benznidazole in trypanosomes. Proc. Natl. Acad. Sci. USA 2008, 105, 5022–5027. [Google Scholar] [CrossRef] [PubMed]
  131. Sokolova, A.Y.; Wyllie, S.; Patterson, S.; Oza, S.L.; Read, K.D.; Fairlamb, A.H. Cross-resistance to nitro drugs and implications for treatment of human African trypanosomiasis. Antimicrob. Agents Chemother. 2010, 54, 2893–2900. [Google Scholar] [CrossRef] [Green Version]
  132. Alsford, S.; Eckert, S.; Baker, N.; Glover, L.; Sanchez-Flores, A.; Leung, K.F.; Turner, D.J.; Field, M.C.; Berriman, M.; Horn, D. High-throughput decoding of antitrypanosomal drug efficacy and resistance. Nature 2012, 482, 232–236. [Google Scholar] [CrossRef] [Green Version]
  133. Thomas, J.A.; Baker, N.; Hutchinson, S.; Dominicus, C.; Trenaman, A.; Glover, L.; Alsford, S.; Horn, D. Insights into antitrypanosomal drug mode-of-action from cytology-based profiling. PLoS Negl. Trop. Dis. 2018, 12, e0006980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Drugs for Neglected Diseases Initiative. Fexinidazole for T. brucei. DNDi—Best Science for the Most Neglected. 16 May 2023. Available online: https://dndi.org/research-development/portfolio/fexinidazole-tb-rhodesiense/ (accessed on 5 May 2023).
  135. Torrico, F.; Gascón, J.; Ortiz, L.; Pinto, J.; Rojas, G.; Palacios, A.; Barreira, F.; Blum, B.; Schijman, A.G.; Vaillant, M.; et al. A Phase 2, Randomized, Multicenter, Placebo-Controlled, Proof-of-Concept Trial of Oral Fexinidazole in Adults With Chronic Indeterminate Chagas Disease. Clin. Infect. Dis. 2023, 76, e1186–e1194. [Google Scholar] [CrossRef] [PubMed]
  136. Dickie, E.A.; Giordani, F.; Gould, M.K.; Mäser, P.; Burri, C.; Mottram, J.C.; Rao, S.P.S.; Barrett, M.P. New Drugs for Human African Trypanosomiasis: A Twenty First Century Success Story. Trop. Med. Infect. Dis. 2020, 5, 29. [Google Scholar] [PubMed] [Green Version]
  137. Waithaka, A.; Clayton, C. Clinically relevant benzoxaboroles inhibit mRNA processing in Trypanosoma brucei. BMC Res. Notes 2022, 15, 371. [Google Scholar] [CrossRef]
  138. Steketee, P.C.; Vincent, I.M.; Achcar, F.; Giordani, F.; Kim, D.-H.; Creek, D.J.; Freund, Y.; Jacobs, R.; Rattigan, K.; Horn, D.; et al. Benzoxaborole treatment perturbs S-adenosyl-L-methionine metabolism in Trypanosoma brucei. PLoS Negl. Trop. Dis. 2018, 12, e0006450. [Google Scholar] [CrossRef] [Green Version]
  139. Wall, R.J.; Rico, E.; Lukac, I.; Zuccotto, F.; Elg, S.; Gilbert, I.H.; Freund, Y.; Alley, M.R.K.; Field, M.C.; Wyllie, S.; et al. Clinical and veterinary trypanocidal benzoxaboroles target CPSF3. Proc. Natl. Acad. Sci. USA 2018, 115, 9616–9621. [Google Scholar] [CrossRef] [Green Version]
  140. Kumeso, V.K.; Kalonji, W.M.; Rembry, S.; Mordt, O.; Ngolo Tete, D.; Prêtre, A.; Delhomme, S.; Kyhi, M.; Camara, M.; Catusse, J.; et al. Efficacy and safety of acoziborole in patients with human African trypanosomiasis caused by Trypanosoma brucei gambiense: A multicentre, open-label, single-arm, phase 2/3 trial. Lancet Infect. Dis. 2023, 23, 463–470. [Google Scholar]
  141. Jacobs, R.T.; Nare, B.; Wring, S.A.; Orr, M.D.; Chen, D.; Sligar, J.M.; Jenks, M.X.; Noe, R.A.; Bowling, T.S.; Mercer, L.T.; et al. SCYX-7158, an Orally-Active Benzoxaborole for the Treatment of Stage 2 Human African Trypanosomiasis. PLoS Negl. Trop. Dis. 2011, 5, e1151. [Google Scholar] [CrossRef] [Green Version]
  142. Dominski, Z.; Yang, X.; Marzluff, W.F. The Polyadenylation Factor CPSF-73 Is Involved in Histone-Pre-mRNA Processing. Cell 2005, 123, 37–48. [Google Scholar] [CrossRef] [Green Version]
  143. Mandel, C.R.; Kaneko, S.; Zhang, H.; Gebauer, D.; Vethantham, V.; Manley, J.L.; Tong, L. Polyadenylation factor CPSF-73 is the pre-mRNA 3′-end-processing endonuclease. Nature 2006, 444, 953–956. [Google Scholar] [CrossRef] [Green Version]
  144. Akama, T.; Zhang, Y.-K.; Freund, Y.R.; Berry, P.; Lee, J.; Easom, E.E.; Jacobs, R.T.; Plattner, J.J.; Witty, M.J.; Peter, R.; et al. Identification of a 4-fluorobenzyl l-valinate amide benzoxaborole (AN11736) as a potential development candidate for the treatment of Animal African Trypanosomiasis (AAT). Bioorg. Med. Chem. Lett. 2018, 28, 6–10. [Google Scholar] [CrossRef] [PubMed]
  145. Burri, C.; Yeramian, P.D.; Allen, J.L.; Merolle, A.; Serge, K.K.; Mpanya, A.; Lutumba, P.; Mesu, V.K.B.K.; Bilenge, C.M.M.; Lubaki, J.-P.F.; et al. Efficacy, Safety, and Dose of Pafuramidine, a New Oral Drug for Treatment of First Stage Sleeping Sickness, in a Phase 2a Clinical Study and Phase 2b Randomized Clinical Studies. PLoS Negl. Trop. Dis. 2016, 10, e0004362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Pohlig, G.; Bernhard, S.C.; Blum, J.; Burri, C.; Mpanya, A.; Lubaki, J.-P.F.; Mpoto, A.M.; Munungu, B.F.; N’tombe, P.M.; Deo, G.K.M.; et al. Efficacy and Safety of Pafuramidine versus Pentamidine Maleate for Treatment of First Stage Sleeping Sickness in a Randomized, Comparator-Controlled, International Phase 3 Clinical Trial. PLoS Negl. Trop. Dis. 2016, 10, e0004363. [Google Scholar] [CrossRef] [Green Version]
  147. Mathis, A.M.; Holman, J.L.; Sturk, L.M.; Ismail, M.A.; Boykin, D.W.; Tidwell, R.R.; Hall, J.E. Accumulation and Intracellular Distribution of Antitrypanosomal Diamidine Compounds DB75 and DB820 in African Trypanosomes. Antimicrob. Agents Chemother. 2006, 50, 2185–2191. [Google Scholar] [CrossRef] [Green Version]
  148. Thuita, J.K.; Wang, M.Z.; Kagira, J.M.; Denton, C.L.; Paine, M.F.; Mdachi, R.E.; Murilla, G.A.; Ching, S.; Boykin, D.W.; Tidwell, R.R.; et al. Pharmacology of DB844, an Orally Active aza Analogue of Pafuramidine, in a Monkey Model of Second Stage Human African Trypanosomiasis. PLoS Negl. Trop. Dis. 2012, 6, e1734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Thuita, J.K.; Wolf, K.K.; Murilla, G.A.; Bridges, A.S.; Boykin, D.W.; Mutuku, J.N.; Liu, Q.; Jones, S.K.; Gem, C.O.; Ching, S.; et al. Chemotherapy of Second Stage Human African Trypanosomiasis: Comparison between the Parenteral Diamidine DB829 and Its Oral Prodrug DB868 in Vervet Monkeys. PLoS Negl. Trop. Dis. 2015, 9, e0003409. [Google Scholar]
  150. De Koning, H.P. The Drugs of Sleeping Sickness: Their Mechanisms of Action and Resistance, and a Brief History. Trop. Med. Infect. Dis. 2020, 5, 14. [Google Scholar] [CrossRef] [Green Version]
  151. Sturk, L.M.; Brock, J.L.; Bagnell, C.R.; Hall, J.E.; Tidwell, R.R. Distribution and quantitation of the anti-trypanosomal diamidine 2,5-bis(4-amidinophenyl)furan (DB75) and its N-methoxy prodrug DB289 in murine brain tissue. Acta Trop. 2004, 91, 131–143. [Google Scholar] [CrossRef]
  152. Rodgers, J.; Jones, A.; Gibaud, S.; Bradley, B.; McCabe, C.; Barrett, M.P.; Gettinby, G.; Kennedy, P.G.E. Melarsoprol Cyclodextrin Inclusion Complexes as Promising Oral Candidates for the Treatment of Human African Trypanosomiasis. PLoS Negl. Trop. Dis. 2011, 5, e1308. [Google Scholar] [CrossRef] [Green Version]
  153. Saokham, P.; Muankaew, C.; Jansook, P.; Loftsson, T. Solubility of Cyclodextrins and Drug/Cyclodextrin Complexes. Molecules 2018, 23, 1161. [Google Scholar] [CrossRef] [Green Version]
  154. Jackson, A.P.; Sanders, M.; Berry, A.; McQuillan, J.; Aslett, M.A.; Quail, M.A.; Chukualim, B.; Capewell, P.; MacLeod, A.; Melville, S.E.; et al. The Genome Sequence of Trypanosoma brucei gambiense, Causative Agent of Chronic Human African Trypanosomiasis. PLoS Negl. Trop. Dis. 2010, 4, e658. [Google Scholar] [CrossRef] [Green Version]
  155. Gibson, W. The origins of the trypanosome genome strains Trypanosoma brucei brucei TREU 927, T. b. gambiense DAL 972, T. vivax Y486 and T. congolense IL3000. Parasit. Vectors 2012, 5, 71. [Google Scholar] [CrossRef] [Green Version]
  156. Kamps, M.P.; Buss, J.E.; Sefton, B.M. Mutation of NH2-terminal glycine of p60src prevents both myristoylation and morphological transformation. Proc. Natl. Acad. Sci. USA 1985, 82, 4625–4628. [Google Scholar] [CrossRef] [PubMed]
  157. Buss, J.E.; Kamps, M.P.; Gould, K.; Sefton, B.M. The absence of myristic acid decreases membrane binding of p60src but does not affect tyrosine protein kinase activity. J. Virol. 1986, 58, 468–474. [Google Scholar] [CrossRef] [PubMed]
  158. Towler, D.A.; Adams, S.P.; Eubanks, S.R.; Towery, D.S.; Jackson-Machelski, E.; Glaser, L.; Gordon, J.I. Purification and characterization of yeast myristoyl CoA:protein N-myristoyltransferase. Proc. Natl. Acad. Sci. USA 1987, 84, 2708–2712. [Google Scholar] [CrossRef]
  159. Zha, J.; Weiler, S.; Oh, K.J.; Wei, M.C.; Korsmeyer, S.J. Posttranslational N-Myristoylation of BID as a Molecular Switch for Targeting Mitochondria and Apoptosis. Science 2000, 290, 1761–1765. [Google Scholar] [CrossRef]
  160. Price, H.P.; Menon, M.R.; Panethymitaki, C.; Goulding, D.; McKean, P.G.; Smith, D.F. Myristoyl-CoA:Protein N-Myristoyltransferase, an Essential Enzyme and Potential Drug Target in Kinetoplastid Parasites. J. Biol. Chem. 2003, 278, 7206–7214. [Google Scholar] [CrossRef] [Green Version]
  161. Price, H.P.; Güther, M.L.S.; Ferguson, M.A.J.; Smith, D.F. Myristoyl-CoA:protein N-myristoyltransferase depletion in trypanosomes causes avirulence and endocytic defects. Mol. Biochem. Parasitol. 2010, 169, 55–58. [Google Scholar] [CrossRef] [PubMed]
  162. Giang, D.K.; Cravatt, B.F. A Second Mammalian N-Myristoyltransferase. J. Biol. Chem. 1998, 273, 6595–6598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Frearson, J.A.; Brand, S.; McElroy, S.P.; Cleghorn, L.A.T.; Smid, O.; Stojanovski, L.; Price, H.P.; Guther, M.L.S.; Torrie, L.S.; Robinson, D.A.; et al. N-myristoyltransferase inhibitors as new leads to treat sleeping sickness. Nature 2010, 464, 728–732. [Google Scholar] [CrossRef] [Green Version]
  164. Hertz-Fowler, C.; Ersfeld, K.; Gull, K. CAP5.5, a life-cycle-regulated, cytoskeleton-associated protein is a member of a novel family of calpain-related proteins in Trypanosoma brucei. Mol. Biochem. Parasitol. 2001, 116, 25–34. [Google Scholar] [CrossRef] [PubMed]
  165. Price, H.P.; Panethymitaki, C.; Goulding, D.; Smith, D.F. Functional analysis of TbARL1, an N-myristoylated Golgi protein essential for viability in bloodstream trypanosomes. J. Cell Sci. 2005, 118, 831–841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Price, H.P.; Stark, M.; Smith, D.F. Trypanosoma brucei ARF1 Plays a Central Role in Endocytosis and Golgi–Lysosome Trafficking. Mol. Biol. Cell 2007, 18, 864–873. [Google Scholar] [CrossRef] [Green Version]
  167. Denny, P.W.; Gokool, S.; Russell, D.G.; Field, M.C.; Smith, D.F. Acylation-dependent Protein Export inLeishmania. J. Biol. Chem. 2000, 275, 11017–11025. [Google Scholar] [CrossRef] [Green Version]
  168. Wingard, J.N.; Ladner, J.; Vanarotti, M.; Fisher, A.J.; Robinson, H.; Buchanan, K.T.; Engman, D.M.; Ames, J.B. Structural insights into membrane targeting by the flagellar calcium-binding protein (FCaBP), a myristoylated and palmitoylated calcium sensor in Trypanosoma cruzi. J. Biol. Chem. 2008, 283, 23388–23396. [Google Scholar] [CrossRef] [Green Version]
  169. Mills, E.; Price, H.P.; Johner, A.; Emerson, J.E.; Smith, D.F. Kinetoplastid PPEF phosphatases: Dual acylated proteins expressed in the endomembrane system of Leishmania. Mol. Biochem. Parasitol. 2007, 152, 22–34. [Google Scholar] [CrossRef]
  170. Brand, S.; Cleghorn, L.A.T.; McElroy, S.P.; Robinson, D.A.; Smith, V.C.; Hallyburton, I.; Harrison, J.R.; Norcross, N.R.; Spinks, D.; Bayliss, T.; et al. Discovery of a Novel Class of Orally Active Trypanocidal N-Myristoyltransferase Inhibitors. J. Med. Chem. 2012, 55, 140–152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Brand, S.; Norcross, N.R.; Thompson, S.; Harrison, J.R.; Smith, V.C.; Robinson, D.A.; Torrie, L.S.; McElroy, S.P.; Hallyburton, I.; Norval, S.; et al. Lead Optimization of a Pyrazole Sulfonamide Series of Trypanosoma brucei N-Myristoyltransferase Inhibitors: Identification and Evaluation of CNS Penetrant Compounds as Potential Treatments for Stage 2 Human African Trypanosomiasis. J. Med. Chem. 2014, 57, 9855–9869. [Google Scholar] [CrossRef]
  172. Bayliss, T.; Robinson, D.A.; Smith, V.C.; Brand, S.; McElroy, S.P.; Torrie, L.S.; Mpamhanga, C.; Norval, S.; Stojanovski, L.; Brenk, R.; et al. Design and Synthesis of Brain Penetrant Trypanocidal N-Myristoyltransferase Inhibitors. J. Med. Chem. 2017, 60, 9790–9806. [Google Scholar] [CrossRef] [Green Version]
  173. Boer, D.R.; Bijlmakers, M.-J. Differential Inhibition of Human and Trypanosome Ubiquitin E1S by TAK-243 Offers Possibilities for Parasite Selective Inhibitors. Sci. Rep. 2019, 9, 16195. [Google Scholar] [CrossRef] [Green Version]
  174. Bijlmakers, M.-J. Ubiquitination and the Proteasome as Drug Targets in Trypanosomatid Diseases. Front. Chem. 2021, 8, 630888. [Google Scholar] [CrossRef] [PubMed]
  175. Pickart, C.M. Mechanisms underlying ubiquitination. Annu. Rev. Biochem. 2001, 70, 503–533. [Google Scholar] [CrossRef] [PubMed]
  176. Hicke, L. Protein regulation by monoubiquitin. Nat. Rev. Mol. Cell Biol. 2001, 2, 195–201. [Google Scholar] [CrossRef] [PubMed]
  177. Peng, J.; Schwartz, D.; Elias, J.E.; Thoreen, C.C.; Cheng, D.; Marsischky, G.; Roelofs, J.; Finley, D.; Gygi, S.P. A proteomics approach to understanding protein ubiquitination. Nat. Biotechnol. 2003, 21, 921–926. [Google Scholar] [CrossRef]
  178. Pickart, C.M.; Fushman, D. Polyubiquitin chains: Polymeric protein signals. Curr. Opin. Chem. Biol. 2004, 8, 610–616. [Google Scholar] [CrossRef]
  179. Hershko, A.; Ciechanover, A. The ubiquitin system. Annu. Rev. Biochem. 1998, 67, 425–479. [Google Scholar] [CrossRef]
  180. Scheffner, M.; Nuber, U.; Huibregtse, J.M. Protein ubiquitination involving an E1–E2–E3 enzyme ubiquitin thioester cascade. Nature 1995, 373, 81–83. [Google Scholar] [CrossRef] [Green Version]
  181. Zhang, L.; Xu, M.; Scotti, E.; Chen, Z.J.; Tontonoz, P. Both K63 and K48 ubiquitin linkages signal lysosomal degradation of the LDL receptor. J. Lipid Res. 2013, 54, 1410–1420. [Google Scholar] [CrossRef] [Green Version]
  182. Nguyen, L.K.; Dobrzyński, M.; Fey, D.; Kholodenko, B.N. Polyubiquitin chain assembly and organization determine the dynamics of protein activation and degradation. Front. Physiol. 2014, 5, 4. [Google Scholar] [CrossRef] [Green Version]
  183. Chung, W.-L.; Leung, K.F.; Carrington, M.; Field, M.C. Ubiquitylation is Required for Degradation of Transmembrane Surface Proteins in Trypanosomes. Traffic 2008, 9, 1681–1697. [Google Scholar] [CrossRef]
  184. Alsford, S.; Turner, D.J.; Obado, S.O.; Sanchez-Flores, A.; Glover, L.; Berriman, M.; Hertz-Fowler, C.; Horn, D. High-throughput phenotyping using parallel sequencing of RNA interference targets in the African trypanosome. Genome Res. 2011, 21, 915–924. [Google Scholar] [CrossRef] [Green Version]
  185. Gupta, I.; Aggarwal, S.; Singh, K.; Yadav, A.; Khan, S. Ubiquitin Proteasome pathway proteins as potential drug targets in parasite Trypanosoma cruzi. Sci. Rep. 2018, 8, 8399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Dewar, C.E.; Oeljeklaus, S.; Mani, J.; Mühlhäuser, W.W.D.; von Känel, C.; Zimmermann, J.; Ochsenreiter, T.; Warscheid, B.; Schneider, A. Mistargeting of aggregation prone mitochondrial proteins activates a nucleus-mediated posttranscriptional quality control pathway in trypanosomes. Nat. Commun. 2022, 13, 3084. [Google Scholar] [CrossRef] [PubMed]
  187. Barghout, S.H.; Schimmer, A.D. E1 Enzymes as Therapeutic Targets in Cancer. Pharmacol. Rev. 2021, 73, 1–56. [Google Scholar] [CrossRef]
  188. Khare, S.; Nagle, A.S.; Biggart, A.; Lai, Y.H.; Liang, F.; Davis, L.C.; Barnes, S.W.; Mathison, C.J.N.; Myburgh, E.; Gao, M.-Y.; et al. Proteasome inhibition for treatment of leishmaniasis, Chagas disease and sleeping sickness. Nature 2016, 537, 229–233. [Google Scholar] [CrossRef] [Green Version]
  189. Wyllie, S.; Brand, S.; Thomas, M.; De Rycker, M.; Chung, C.-W.; Pena, I.; Bingham, R.P.; Bueren-Calabuig, J.A.; Cantizani, J.; Cebrian, D.; et al. Preclinical candidate for the treatment of visceral leishmaniasis that acts through proteasome inhibition. Proc. Natl. Acad. Sci. USA 2019, 116, 9318–9323. [Google Scholar] [CrossRef] [Green Version]
  190. Nagle, A.; Biggart, A.; Be, C.; Srinivas, H.; Hein, A.; Caridha, D.; Sciotti, R.J.; Pybus, B.; Kreishman-Deitrick, M.; Bursulaya, B.; et al. Discovery and Characterization of Clinical Candidate LXE408 as a Kinetoplastid-Selective Proteasome Inhibitor for the Treatment of Leishmaniases. J. Med. Chem. 2020, 63, 10773–10781. [Google Scholar] [CrossRef]
  191. Pfarr, K.M.; Krome, A.K.; Al-Obaidi, I.; Batchelor, H.; Vaillant, M.; Hoerauf, A.; Opoku, N.O.; Kuesel, A.C. The pipeline for drugs for control and elimination of neglected tropical diseases: 1. Anti-infective drugs for regulatory registration. Parasit. Vectors 2023, 16, 82. [Google Scholar]
  192. Oberholzer, M.; Marti, G.; Baresic, M.; Kunz, S.; Hemphill, A.; Seebeck, T. The Trypanosoma brucei cAMP phosphodiesterases TbrPDEBl and TbrPDEB2: Flagellar enzymes that are essential for parasite virulence. FASEB J. 2007, 21, 720–731. [Google Scholar] [CrossRef] [PubMed]
  193. Luginbuehl, E.; Ryter, D.; Schranz-Zumkehr, J.; Oberholzer, M.; Kunz, S.; Seebeck, T. The N Terminus of Phosphodiesterase TbrPDEB1 of Trypanosoma brucei Contains the Signal for Integration into the Flagellar Skeleton. Eukaryot. Cell 2010, 9, 1466–1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. De Koning, H.P.; Gould, M.K.; Sterk, G.J.; Tenor, H.; Kunz, S.; Luginbuehl, E.; Seebeck, T. Pharmacological Validation of Trypanosoma brucei Phosphodiesterases as Novel Drug Targets. J. Infect. Dis. 2012, 206, 229–237. [Google Scholar] [CrossRef]
  195. Seebeck, T.; Sterk, G.J.; Ke, H. Phosphodiesterase inhibitors as a new generation of antiprotozoan drugs: Exploiting the benefit of enzymes that are highly conserved between host and parasite. Future Med. Chem. 2011, 3, 1289–1306. [Google Scholar]
  196. Gould, M.K.; Bachmaier, S.; Ali, J.A.M.; Alsford, S.; Tagoe, D.N.A.; Munday, J.C.; Schnaufer, A.C.; Horn, D.; Boshart, M.; De Koning, H.P. Cyclic AMP Effectors in African Trypanosomes Revealed by Genome-Scale RNA Interference Library Screening for Resistance to the Phosphodiesterase Inhibitor CpdA. Antimicrob. Agents Chemother. 2013, 57, 4882–4893. [Google Scholar] [CrossRef] [Green Version]
  197. Bachmaier, S.; Gould, M.K.; Polatoglou, E.; Omelianczyk, R.; Brennand, A.E.; Aloraini, M.A.; Munday, J.C.; Horn, D.; Boshart, M.; De Koning, H.P. Novel kinetoplastid-specific cAMP binding proteins identified by RNAi screening for cAMP resistance in Trypanosoma brucei. Front. Cell. Infect. Microbiol. 2023, 13, 1204707. [Google Scholar] [CrossRef]
  198. Van Der Mey, M.; Hatzelmann, A.; Van Der Laan, I.J.; Sterk, G.J.; Thibaut, U.; Timmerman, H. Novel Selective PDE4 Inhibitors. 1. Synthesis, Structure−Activity Relationships, and Molecular Modeling of 4-(3,4-Dimethoxyphenyl)-2H-phthalazin-1-ones and Analogues. J. Med. Chem. 2001, 44, 2511–2522. [Google Scholar] [CrossRef] [PubMed]
  199. Van Der Mey, M.; Hatzelmann, A.; Van Klink, G.P.M.; Van Der Laan, I.J.; Sterk, G.J.; Thibaut, U.; Ulrich, W.R.; Timmerman, H. Novel Selective PDE4 Inhibitors. 2. Synthesis and Structure−Activity Relationships of 4-Aryl-Substituted cis-Tetra- and cis-Hexahydrophthalazinones. J. Med. Chem. 2001, 44, 2523–2535. [Google Scholar] [CrossRef] [PubMed]
  200. Souness, J.E.; Aldous, D.; Sargent, C. Immunosuppressive and anti-inflammatory effects of cyclic AMP phosphodiesterase (PDE) type 4 inhibitors. Immunopharmacology 2000, 47, 127–162. [Google Scholar] [CrossRef]
  201. Teixeira, M.M.; Gristwood, R.W.; Cooper, N.; Hellewell, P.G. Phosphodiesterase (PDE)4 inhibitors: Anti-inflammatory drugs of the future? Trends Pharmacol. Sci. 1997, 18, 164–170. [Google Scholar] [CrossRef]
  202. Veerman, J.; Van Den Bergh, T.; Orrling, K.M.; Jansen, C.; Cos, P.; Maes, L.; Chatelain, E.; Ioset, J.-R.; Edink, E.E.; Tenor, H.; et al. Synthesis and evaluation of analogs of the phenylpyridazinone NPD-001 as potent trypanosomal TbrPDEB1 phosphodiesterase inhibitors and in vitro trypanocidals. Bioorg. Med. Chem. 2016, 24, 1573–1581. [Google Scholar] [CrossRef] [PubMed]
  203. De Heuvel, E.; Singh, A.K.; Boronat, P.; Kooistra, A.J.; Van Der Meer, T.; Sadek, P.; Blaazer, A.R.; Shaner, N.C.; Bindels, D.S.; Caljon, G.; et al. Alkynamide phthalazinones as a new class of TbrPDEB1 inhibitors (Part 2). Bioorg. Med. Chem. 2019, 27, 4013–4029. [Google Scholar] [CrossRef] [PubMed]
  204. Fairlamb, A.H.; Cerami, A. Identification of a novel, thiol-containing co-factor essential for glutathione reductase enzyme activity in trypanosomatids. Mol. Biochem. Parasitol. 1985, 14, 187–198. [Google Scholar] [CrossRef] [PubMed]
  205. Battista, T.; Colotti, G.; Ilari, A.; Fiorillo, A. Targeting Trypanothione Reductase, a Key Enzyme in the Redox Trypanosomatid Metabolism, to Develop New Drugs against Leishmaniasis and Trypanosomiases. Molecules 2020, 25, 1924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Shames, S.L.; Fairlamb, A.H.; Cerami, A.; Walsh, C.T. Purification and characterization of trypanothione reductase from Crithidia fasciculata, a new member of the family of disulfide-containing flavoprotein reductases. Biochemistry 1986, 25, 3519–3526. [Google Scholar] [CrossRef] [PubMed]
  207. Krieger, S.; Schwarz, W.; Ariyanayagam, M.R.; Fairlamb, A.H.; Krauth-Siegel, R.L.; Clayton, C. Trypanosomes lacking trypanothione reductase are avirulent and show increased sensitivity to oxidative stress: Trypanosomes lacking trypanothione reductase. Mol. Microbiol. 2002, 35, 542–552. [Google Scholar]
  208. Stoll, V.S.; Simpson, S.J.; Krauth-Siegel, R.L.; Walsh, C.T.; Pai, E.F. Glutathione Reductase Turned into Trypanothione Reductase: Structural Analysis of an Engineered Change in Substrate Specificity. Biochemistry 1997, 36, 6437–6447. [Google Scholar] [CrossRef]
  209. Zhang, Y.; Bond, C.S.; Bailey, S.; Cunningham, M.L.; Fairlamb, A.H.; Hunter, W.N. The crystal structure of trypanothione reductase from the human pathogen Trypanosoma cruzi at 2.3 Å resolution: Trypanothione reductase structure. Protein Sci. 1996, 5, 52–61. [Google Scholar] [CrossRef] [Green Version]
  210. Beig, M.; Oellien, F.; Garoff, L.; Noack, S.; Krauth-Siegel, R.L.; Selzer, P.M. Trypanothione Reductase: A Target Protein for a Combined In Vitro and In Silico Screening Approach. PLoS Negl. Trop. Dis. 2015, 9, e0003773. [Google Scholar] [CrossRef] [Green Version]
  211. Tovar, J.; Cunningham, M.L.; Smith, A.C.; Croft, S.L.; Fairlamb, A.H. Down-regulation of Leishmania donovani trypanothione reductase by heterologous expression of a trans-dominant mutant homologue: Effect on parasite intracellular survival. Proc. Natl. Acad. Sci. USA 1998, 95, 5311–5316. [Google Scholar] [CrossRef]
  212. Turcano, L.; Battista, T.; De Haro, E.T.; Missineo, A.; Alli, C.; Paonessa, G.; Colotti, G.; Harper, S.; Fiorillo, A.; Ilari, A.; et al. Spiro-containing derivatives show antiparasitic activity against Trypanosoma brucei through inhibition of the trypanothione reductase enzyme. PLoS Negl. Trop. Dis. 2020, 14, e0008339. [Google Scholar] [CrossRef]
  213. Fairlamb, A.H.; Blackburn, P.; Ulrich, P.; Chait, B.T.; Cerami, A. Trypanothione: A Novel Bis(glutathionyl)spermidine Cofactor for Glutathione Reductase in Trypanosomatids. Science 1985, 227, 1485–1487. [Google Scholar] [CrossRef]
  214. Oza, S.L.; Ariyanayagam, M.R.; Aitcheson, N.; Fairlamb, A.H. Properties of trypanothione synthetase from Trypanosoma brucei. Mol. Biochem. Parasitol. 2003, 131, 25–33. [Google Scholar] [CrossRef]
  215. Comini, M.A.; Guerrero, S.A.; Haile, S.; Menge, U.; Lünsdorf, H.; Flohé, L. Valdiation of Trypanosoma brucei trypanothione synthetase as drug target. Free Radic. Biol. Med. 2004, 36, 1289–1302. [Google Scholar] [CrossRef]
  216. Mesías, A.C.; Sasoni, N.; Arias, D.G.; Pérez Brandán, C.; Orban, O.C.F.; Kunick, C.; Robello, C.; Comini, M.A.; Garg, N.J.; Zago, M.P. Trypanothione synthetase confers growth, survival advantage and resistance to anti-protozoal drugs in Trypanosoma cruzi. Free Radic. Biol. Med. 2019, 130, 23–34. [Google Scholar] [CrossRef]
  217. Benítez, D.; Franco, J.; Sardi, F.; Leyva, A.; Durán, R.; Choi, G.; Yang, G.; Kim, T.; Kim, N.; Heo, J.; et al. Drug-like molecules with anti-trypanothione synthetase activity identified by high throughput screening. J. Enzyme Inhib. Med. Chem. 2022, 37, 912–929. [Google Scholar] [CrossRef]
  218. Fairlamb, A.H.; Henderson, G.B.; Bacchi, C.J.; Cerami, A. In vivo effects of difluoromethylornithine on trypanothione and polyamine levels in bloodstream forms of Trypanosoma brucei. Mol. Biochem. Parasitol. 1987, 24, 185–191. [Google Scholar] [CrossRef]
  219. Yun, O.; Priotto, G.; Tong, J.; Flevaud, L.; Chappuis, F. NECT is next: Implementing the new drug combination therapy for Trypanosoma brucei gambiense sleeping sickness. PLoS Negl. Trop. Dis. 2010, 4, e720. [Google Scholar] [CrossRef] [Green Version]
  220. Heby, O.; Persson, L.; Rentala, M. Targeting the polyamine biosynthetic enzymes: A promising approach to therapy of African sleeping sickness, Chagas’ disease, and leishmaniasis. Amino Acids 2007, 33, 359–366. [Google Scholar] [CrossRef] [PubMed]
  221. Willert, E.K.; Phillips, M.A. Regulated Expression of an Essential Allosteric Activator of Polyamine Biosynthesis in African Trypanosomes. PLoS Pathog. 2008, 4, e1000183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Taylor, M.C.; Kaur, H.; Blessington, B.; Kelly, J.M.; Wilkinson, S.R. Validation of spermidine synthase as a drug target in African trypanosomes. Biochem. J. 2008, 409, 563–569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Pegg, A.E. S-Adenosylmethionine decarboxylase. Essays Biochem. 2009, 46, 25–46. [Google Scholar] [PubMed] [Green Version]
  224. Volkov, O.A.; Cosner, C.C.; Brockway, A.J.; Kramer, M.; Booker, M.; Zhong, S.; Ketcherside, A.; Wei, S.; Longgood, J.; McCoy, M.; et al. Identification of Trypanosoma brucei AdoMetDC Inhibitors Using a High-Throughput Mass Spectrometry-Based Assay. ACS Infect. Dis. 2017, 3, 512–526. [Google Scholar] [CrossRef]
  225. Ekstrom, J.L.; Tolbert, W.D.; Xiong, H.; Pegg, A.E.; Ealick, S.E. Structure of a Human S-Adenosylmethionine Decarboxylase Self-Processing Ester Intermediate and Mechanism of Putrescine Stimulation of Processing As Revealed by the H243A Mutant. Biochemistry 2001, 40, 9495–9504. [Google Scholar] [CrossRef]
  226. Tolbert, W.D.; Zhang, Y.; Cottet, S.E.; Bennett, E.M.; Ekstrom, J.L.; Pegg, A.E.; Ealick, S.E. Mechanism of Human S-Adenosylmethionine Decarboxylase Proenzyme Processing As Revealed by the Structure of the S68A Mutant. Biochemistry 2003, 42, 2386–2395. [Google Scholar] [CrossRef] [PubMed]
  227. Bale, S.; Ealick, S.E. Structural biology of S-adenosylmethionine decarboxylase. Amino Acids 2010, 38, 451–460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Willert, E.K.; Fitzpatrick, R.; Phillips, M.A. Allosteric regulation of an essential trypanosome polyamine biosynthetic enzyme by a catalytically dead homolog. Proc. Natl. Acad. Sci. USA 2007, 104, 8275–8280. [Google Scholar] [CrossRef]
  229. Velez, N.; Brautigam, C.A.; Phillips, M.A. Trypanosoma brucei S-Adenosylmethionine Decarboxylase N Terminus Is Essential for Allosteric Activation by the Regulatory Subunit Prozyme. J. Biol. Chem. 2013, 288, 5232–5240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Rusché, L.N.; Huang, C.E.; Piller, K.J.; Hemann, M.; Wirtz, E.; Sollner-Webb, B. The two RNA ligases of the Trypanosoma brucei RNA editing complex: Cloning the essential band IV gene and identifying the band V gene. Mol. Cell. Biol. 2001, 21, 979–989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Schnaufer, A.; Panigrahi, A.K.; Panicucci, B.; Igo, R.P.; Wirtz, E.; Salavati, R.; Stuart, K. An RNA ligase essential for RNA editing and survival of the bloodstream form of Trypanosoma brucei. Science 2001, 291, 2159–2162. [Google Scholar] [CrossRef]
  232. Amaro, R.E.; Schnaufer, A.; Interthal, H.; Hol, W.; Stuart, K.D.; McCammon, J.A. Discovery of drug-like inhibitors of an essential RNA-editing ligase in Trypanosoma brucei. Proc. Natl. Acad. Sci. USA 2008, 105, 17278–17283. [Google Scholar] [CrossRef]
  233. Bello, A.R.; Nare, B.; Freedman, D.; Hardy, L.; Beverley, S.M. PTR1: A reductase mediating salvage of oxidized pteridines and methotrexate resistance in the protozoan parasite Leishmania major. Proc. Natl. Acad. Sci. USA 1994, 91, 11442–11446. [Google Scholar] [CrossRef]
  234. Sienkiewicz, N.; Ong, H.B.; Fairlamb, A.H. Trypanosoma brucei pteridine reductase 1 is essential for survival in vitro and for virulence in mice. Mol. Microbiol. 2010, 77, 658–671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Robello, C.; Navarro, P.; Castanys, S.; Gamarro, F. A pteridine reductase gene ptr1 contiguous to a P-glycoprotein confers resistance to antifolates in Trypanosoma cruzi. Mol. Biochem. Parasitol. 1997, 90, 525–535. [Google Scholar] [CrossRef] [PubMed]
  236. Cavazzuti, A.; Paglietti, G.; Hunter, W.N.; Gamarro, F.; Piras, S.; Loriga, M.; Allecca, S.; Corona, P.; McLuskey, K.; Tulloch, L.; et al. Discovery of potent pteridine reductase inhibitors to guide antiparasite drug development. Proc. Natl. Acad. Sci. USA 2008, 105, 1448–1453. [Google Scholar] [CrossRef] [PubMed]
  237. Tulloch, L.B.; Martini, V.P.; Iulek, J.; Huggan, J.K.; Lee, J.H.; Gibson, C.L.; Smith, T.K.; Suckling, C.J.; Hunter, W.N. Structure-Based Design of Pteridine Reductase Inhibitors Targeting African Sleeping Sickness and the Leishmaniases. J. Med. Chem. 2010, 53, 221–229. [Google Scholar] [CrossRef] [PubMed]
  238. Kimuda, M.P.; Laming, D.; Hoppe, H.C.; Tastan Bishop, Ö. Identification of Novel Potential Inhibitors of Pteridine Reductase 1 in Trypanosoma brucei via Computational Structure-Based Approaches and in Vitro Inhibition Assays. Molecules 2019, 24, 142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Landi, G.; Linciano, P.; Borsari, C.; Bertolacini, C.P.; Moraes, C.B.; Cordeiro-da-Silva, A.; Gul, S.; Witt, G.; Kuzikov, M.; Costi, M.P.; et al. Structural Insights into the Development of Cycloguanil Derivatives as Trypanosoma brucei Pteridine-Reductase-1 Inhibitors. ACS Infect. Dis. 2019, 5, 1105–1114. [Google Scholar] [CrossRef]
  240. Chow, T.Y.-K.; Alaoui-Jamali, M.A.; Yeh, C.; Yuen, L.; Griller, D. The DNA double-stranded break repair protein endo-exonuclease as a therapeutic target for cancer. Mol. Cancer Ther. 2004, 3, 911–919. [Google Scholar] [CrossRef]
  241. Jung, H.-J.; Suh, S.-I.; Suh, M.-H.; Baek, W.-K.; Park, J.-W. Pentamidine reduces expression of hypoxia-inducible factor-1α in DU145 and MDA-MB-231 cancer cells. Cancer Lett. 2011, 303, 39–46. [Google Scholar] [CrossRef]
  242. Wu, Y.; Zhang, Z.; Kou, Z. Pentamidine Inhibits Ovarian Cancer Cell Proliferation and Migration by Maintaining Stability of PTEN in vitro. Drug Des. Devel. Ther. 2021, 15, 2857–2868. [Google Scholar] [CrossRef]
  243. Martínez-Flórez, A.; Galizzi, M.; Izquierdo, L.; Bustamante, J.M.; Rodriguez, A.; Rodriguez, F.; Rodríguez-Cortés, A.; Alberola, J. Repurposing bioenergetic modulators against protozoan parasites responsible for tropical diseases. Int. J. Parasitol. Drugs Drug Resist. 2020, 14, 17–27. [Google Scholar] [CrossRef] [PubMed]
  244. Rohlenova, K.; Sachaphibulkij, K.; Stursa, J.; Bezawork-Geleta, A.; Blecha, J.; Endaya, B.; Werner, L.; Cerny, J.; Zobalova, R.; Goodwin, J.; et al. Selective Disruption of Respiratory Supercomplexes as a New Strategy to Suppress Her2 high Breast Cancer. Antioxid. Redox Signal. 2017, 26, 84–103. [Google Scholar] [CrossRef] [Green Version]
  245. Arbon, D.; Ženíšková, K.; Šubrtová, K.; Mach, J.; Štursa, J.; Machado, M.; Zahedifard, F.; Leštinová, T.; Hierro-Yap, C.; Neuzil, J.; et al. Repurposing of MitoTam: Novel Anti-Cancer Drug Candidate Exhibits Potent Activity against Major Protozoan and Fungal Pathogens. Antimicrob. Agents Chemother. 2022, 66, e00727-22. [Google Scholar] [CrossRef]
  246. Burger, A.; Ludewig, M.H.; Boshoff, A. Investigating the Chaperone Properties of a Novel Heat Shock Protein, Hsp70.c, from Trypanosoma brucei. J. Parasitol. Res. 2014, 2014, 172582. [Google Scholar] [CrossRef] [Green Version]
  247. Ludewig, M.H.; Boshoff, A.; Horn, D.; Blatch, G.L. Trypanosoma brucei J protein 2 is a stress inducible and essential Hsp40. Int. J. Biochem. Cell Biol. 2015, 60, 93–98. [Google Scholar] [CrossRef] [PubMed]
  248. Bentley, S.J.; Boshoff, A. Trypanosoma brucei J-Protein 2 Functionally Co-Operates with the Cytosolic Hsp70 and Hsp70.4 Proteins. Int. J. Mol. Sci. 2019, 20, 5843. [Google Scholar] [CrossRef] [PubMed]
  249. Meyer, K.J.; Shapiro, T.A. Cytosolic and Mitochondrial Hsp90 in Cytokinesis, Mitochondrial DNA Replication, and Drug Action in Trypanosoma brucei. Antimicrob. Agents Chemother. 2021, 65, e00632-21. [Google Scholar] [CrossRef]
  250. Jamabo, M.; Bentley, S.J.; Macucule-Tinga, P.; Tembo, P.; Edkins, A.L.; Boshoff, A. In silico analysis of the HSP90 chaperone system from the African trypanosome, Trypanosoma brucei. Front. Mol. Biosci. 2022, 9, 947078. [Google Scholar] [CrossRef]
  251. Louw, C.A.; Ludewig, M.H.; Mayer, J.; Blatch, G.L. The Hsp70 chaperones of the Tritryps are characterized by unusual features and novel members. Parasitol. Int. 2010, 59, 497–505. [Google Scholar] [CrossRef]
  252. Bentley, S.J.; Jamabo, M.; Boshoff, A. The Hsp70/J-protein machinery of the African trypanosome, Trypanosoma brucei. Cell Stress Chaperones 2019, 24, 125–148. [Google Scholar] [CrossRef] [PubMed]
  253. Andreassend, S.K.; Bentley, S.J.; Blatch, G.L.; Boshoff, A.; Keyzers, R.A. Screening for Small Molecule Modulators of Trypanosoma brucei Hsp70 Chaperone Activity Based upon Alcyonarian Coral-Derived Natural Products. Mar. Drugs 2020, 18, E81. [Google Scholar] [CrossRef] [Green Version]
  254. Burger, A.; Macucule-Tinga, P.; Bentley, S.J.; Ludewig, M.H.; Mhlongo, N.N.; Shonhai, A.; Boshoff, A. Characterization of an Atypical Trypanosoma brucei Hsp70 Demonstrates Its Cytosolic-Nuclear Localization and Modulation by Quercetin and Methylene Blue. Int. J. Mol. Sci. 2021, 22, 6776. [Google Scholar] [CrossRef] [PubMed]
  255. Odunuga, O.O.; Longshaw, V.M.; Blatch, G.L. Hop: More than an Hsp70/Hsp90 adaptor protein. BioEssays 2004, 26, 1058–1068. [Google Scholar] [CrossRef] [PubMed]
  256. Assimon, V.; Gillies, A.; Rauch, J.; Gestwicki, J. Hsp70 Protein Complexes as Drug Targets. Curr. Pharm. Des. 2013, 19, 404–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Miyata, Y. Hsp90 inhibitor geldanamycin and its derivatives as novel cancer chemotherapeutic agents. Curr. Pharm. Des. 2005, 11, 1131–1138. [Google Scholar] [CrossRef]
  258. Mahalingam, D.; Swords, R.; Carew, J.S.; Nawrocki, S.T.; Bhalla, K.; Giles, F.J. Targeting HSP90 for cancer therapy. Br. J. Cancer 2009, 100, 1523–1529. [Google Scholar] [CrossRef] [Green Version]
  259. Butler, L.M.; Ferraldeschi, R.; Armstrong, H.K.; Centenera, M.M.; Workman, P. Maximizing the Therapeutic Potential of HSP90 Inhibitors. Mol. Cancer Res. 2015, 13, 1445–1451. [Google Scholar] [CrossRef] [Green Version]
  260. Sanchez, J.; Carter, T.R.; Cohen, M.S.; Blagg, B.S.J. Old and New Approaches to Target the Hsp90 Chaperone. Curr. Cancer Drug Targets 2020, 20, 253–270. [Google Scholar] [CrossRef]
  261. Pizarro, J.C.; Hills, T.; Senisterra, G.; Wernimont, A.K.; Mackenzie, C.; Norcross, N.R.; Ferguson, M.A.J.; Wyatt, P.G.; Gilbert, I.H.; Hui, R. Exploring the Trypanosoma brucei Hsp83 Potential as a Target for Structure Guided Drug Design. PLoS Negl. Trop. Dis. 2013, 7, e2492. [Google Scholar] [CrossRef] [Green Version]
  262. Jones, C.; Anderson, S.; Singha, U.K.; Chaudhuri, M. Protein phosphatase 5 is required for Hsp90 function during proteotoxic stresses in Trypanosoma brucei. Parasitol. Res. 2008, 102, 835–844. [Google Scholar] [CrossRef]
  263. Meyer, K.J.; Shapiro, T.A. Potent antitrypanosomal activities of heat shock protein 90 inhibitors in vitro and in vivo. J. Infect. Dis. 2013, 208, 489–499. [Google Scholar] [CrossRef] [Green Version]
  264. Pacey, S.; Wilson, R.H.; Walton, M.; Eatock, M.M.; Hardcastle, A.; Zetterlund, A.; Arkenau, H.-T.; Moreno-Farre, J.; Banerji, U.; Roels, B.; et al. A Phase I Study of the Heat Shock Protein 90 Inhibitor Alvespimycin (17-DMAG) Given Intravenously to Patients with Advanced Solid Tumors. Clin. Cancer Res. 2011, 17, 1561–1570. [Google Scholar] [CrossRef] [Green Version]
  265. Anderson, S.; Jones, C.; Saha, L.; Chaudhuri, M. Functional characterization of the serine/threonine protein phosphatase 5 from Trypanosoma brucei. J. Parasitol. 2006, 92, 1152–1161. [Google Scholar] [CrossRef] [PubMed]
  266. Ogbunude, P.O.J.; Ikediobi, C.O. Comparative aspects of purine metabolism in some African trypanosomes. Mol. Biochem. Parasitol. 1983, 9, 279–287. [Google Scholar] [CrossRef] [PubMed]
  267. Hammond, D.J.; Gutteridge, W.E. Purine and pyrimidine metabolism in the trypanosomatidae. Mol. Biochem. Parasitol. 1984, 13, 243–261. [Google Scholar] [CrossRef] [PubMed]
  268. Hofer, A. Targeting the nucleotide metabolism of Trypanosoma brucei and other trypanosomatids. FEMS Microbiol. Rev. 2023, 47, fuad020. [Google Scholar] [CrossRef] [PubMed]
  269. Campagnaro, G.D. Purine Transporters as Efficient Carriers for Anti-kinetoplastid Molecules: 3′-Deoxytubercidin versus Trypanosomes. ACS Infect. Dis. 2022, 8, 1727–1730. [Google Scholar] [CrossRef] [PubMed]
  270. De Koning, H.P.; Jarvis, S.M. Adenosine Transporters in Bloodstream Forms of Trypanosoma brucei brucei: Substrate Recognition Motifs and Affinity for Trypanocidal Drugs. Mol. Pharmacol. 1999, 56, 1162–1170. [Google Scholar] [CrossRef]
  271. Mäser, P.; Sütterlin, C.; Kralli, A.; Kaminsky, R. A Nucleoside Transporter from Trypanosoma brucei Involved in Drug Resistance. Science 1999, 285, 242–244. [Google Scholar] [CrossRef]
  272. Li, J.Y.; Boado, R.J.; Pardridge, W.M. Cloned Blood–Brain Barrier Adenosine Transporter is Identical to the Rat Concentrative Na + Nucleoside Cotransporter CNT2. J. Cereb. Blood Flow Metab. 2001, 21, 929–936. [Google Scholar] [CrossRef] [Green Version]
  273. Drew, M.E.; Morris, J.C.; Wang, Z.; Wells, L.; Sanchez, M.; Landfear, S.M.; Englund, P.T. The Adenosine Analog Tubercidin Inhibits Glycolysis in Trypanosoma brucei as Revealed by an RNA Interference Library. J. Biol. Chem. 2003, 278, 46596–46600. [Google Scholar] [CrossRef] [Green Version]
  274. Vodnala, S.K.; Ferella, M.; Lundén-Miguel, H.; Betha, E.; Van Reet, N.; Amin, D.N.; Öberg, B.; Andersson, B.; Kristensson, K.; Wigzell, H.; et al. Preclinical Assessment of the Treatment of Second-Stage African Trypanosomiasis with Cordycepin and Deoxycoformycin. PLoS Negl. Trop. Dis. 2009, 3, e495. [Google Scholar] [CrossRef] [PubMed]
  275. Kaplinsky, C.; Yeger, H.; Estrov, Z.; Barankiewicz, J.; Pawlin, G.; Freedman, M.H.; Cohen, A. Selective protection of tubercidin toxicity by nitrobenzyl thioinosine in normal tissues but not in human neuroblastoma cells. Cancer Chemother. Pharmacol. 1986, 17, 264–268. [Google Scholar] [CrossRef] [PubMed]
  276. Hulpia, F.; Mabille, D.; Campagnaro, G.D.; Schumann, G.; Maes, L.; Roditi, I.; Hofer, A.; De Koning, H.P.; Caljon, G.; Van Calenbergh, S. Combining tubercidin and cordycepin scaffolds results in highly active candidates to treat late-stage sleeping sickness. Nat. Commun. 2019, 10, 5564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Hulpia, F.; Bouton, J.; Campagnaro, G.D.; Alfayez, I.A.; Mabille, D.; Maes, L.; De Koning, H.P.; Caljon, G.; Van Calenbergh, S. C6–O-alkylated 7-deazainosine nucleoside analogues: Discovery of potent and selective anti-sleeping sickness agents. Eur. J. Med. Chem. 2020, 188, 112018. [Google Scholar] [CrossRef] [PubMed]
  278. Rottenberg, M.E.; Masocha, W.; Ferella, M.; Petitto-Assis, F.; Goto, H.; Kristensson, K.; McCaffrey, R.; Wigzell, H. Treatment of African Trypanosomiasis with Cordycepin and Adenosine Deaminase Inhibitors in a Mouse Model. J. Infect. Dis. 2005, 192, 1658–1665. [Google Scholar] [CrossRef] [Green Version]
  279. Ziegelbauer, K.; Overath, P. Identification of invariant surface glycoproteins in the bloodstream stage of Trypanosoma brucei. J. Biol. Chem. 1992, 267, 10791–10796. [Google Scholar] [CrossRef]
  280. Ziegelbauer, K.; Overath, P. Organization of two invariant surface glycoproteins in the surface coat of Trypanosoma brucei. Infect. Immun. 1993, 61, 4540–4545. [Google Scholar] [CrossRef]
  281. Salmon, D.; Geuskens, M.; Hanocq, F.; Hanocq-Quertier, J.; Nolan, D.; Ruben, L.; Pays, E. A novel heterodimeric transferrin receptor encoded by a pair of VSG expression site-associated genes in T. brucei. Cell 1994, 78, 75–86. [Google Scholar] [CrossRef]
  282. Nolan, D.P.; Jackson, D.G.; Biggs, M.J.; Brabazon, E.D.; Pays, A.; Van Laethem, F.; Paturiaux-Hanocq, F.; Elliot, J.F.; Voorheis, H.P.; Pays, E. Characterization of a Novel Alanine-rich Protein Located in Surface Microdomains in Trypanosoma brucei. J. Biol. Chem. 2000, 275, 4072–4080. [Google Scholar] [CrossRef] [Green Version]
  283. Mehlert, A.; Wormald, M.R.; Ferguson, M.A.J. Modeling of the N-Glycosylated Transferrin Receptor Suggests How Transferrin Binding Can Occur within the Surface Coat of Trypanosoma brucei. PLoS Pathog. 2012, 8, e1002618. [Google Scholar] [CrossRef] [Green Version]
  284. Horn, D. Antigenic variation in African trypanosomes. Mol. Biochem. Parasitol. 2014, 195, 123–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Radwanska, M.; Guirnalda, P.; De Trez, C.; Ryffel, B.; Black, S.; Magez, S. Trypanosomiasis-Induced B Cell Apoptosis Results in Loss of Protective Anti-Parasite Antibody Responses and Abolishment of Vaccine-Induced Memory Responses. PLoS Pathog. 2008, 4, e1000078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Lança, A.S.C.; de Sousa, K.P.; Atouguia, J.; Prazeres, D.M.F.; Monteiro, G.A.; Silva, M.S. Trypanosoma brucei: Immunisation with plasmid DNA encoding invariant surface glycoprotein gene is able to induce partial protection in experimental African trypanosomiasis. Exp. Parasitol. 2011, 127, 18–24. [Google Scholar]
  287. Geiger, A.; Hirtz, C.; Bécue, T.; Bellard, E.; Centeno, D.; Gargani, D.; Rossignol, M.; Cuny, G.; Peltier, J.-B. Exocytosis and protein secretion in Trypanosoma. BMC Microbiol. 2010, 10, 20. [Google Scholar] [CrossRef] [Green Version]
  288. Magez, S.; Li, Z.; Nguyen, H.T.T.; Pinto Torres, J.E.; Van Wielendaele, P.; Radwanska, M.; Began, J.; Zoll, S.; Sterckx, Y.G.-J. The History of Anti-Trypanosome Vaccine Development Shows That Highly Immunogenic and Exposed Pathogen-Derived Antigens Are Not Necessarily Good Target Candidates: Enolase and ISG75 as Examples. Pathogens 2021, 10, 1050. [Google Scholar] [CrossRef]
  289. Franco, J.R.; Cecchi, G.; Paone, M.; Diarra, A.; Grout, L.; Kadima Ebeja, A.; Simarro, P.P.; Zhao, W.; Argaw, D. The elimination of human African trypanosomiasis: Achievements in relation to WHO road map targets for 2020. PLoS Negl. Trop. Dis. 2022, 16, e0010047. [Google Scholar] [CrossRef] [PubMed]
  290. Borlase, A.; Le Rutte, E.A.; Castaño, S.; Blok, D.J.; Toor, J.; Giardina, F.; Davis, E.L.; Aliee, M.; Anderson, R.M.; Ayabina, D.; et al. Evaluating and mitigating the potential indirect effect of COVID-19 on control programmes for seven neglected tropical diseases: A modelling study. Lancet Glob. Health 2022, 10, e1600–e1611. [Google Scholar]
  291. The Independent. Uganda Eliminates Sleeping Sickness as Public Health Problem: MOH. The Independent Uganda. 22 October 2022. Available online: https://www.independent.co.ug/uganda-eliminates-sleeping-sickness-as-public-health-problem-moh/ (accessed on 3 November 2022).
  292. Büscher, P.; Bart, J.-M.; Boelaert, M.; Bucheton, B.; Cecchi, G.; Chitnis, N.; Courtin, D.; Figueiredo, L.M.; Franco, J.-R.; Grébaut, P.; et al. Do Cryptic Reservoirs Threaten Gambiense-Sleeping Sickness Elimination? Trends Parasitol. 2018, 34, 197–207. [Google Scholar] [CrossRef] [Green Version]
  293. Barrett, M.P.; Croft, S.L. Management of trypanosomiasis and leishmaniasis. Br. Med. Bull. 2012, 104, 175–196. [Google Scholar] [CrossRef] [Green Version]
  294. Abaza, S. Recent advances in identification of potential drug targets and development of novel drugs in parasitic diseases. Part III: Helminths. Parasitol. United J. 2022, 15, 126–143. [Google Scholar] [CrossRef]
  295. Ross, N.T.; Lohmann, F.; Carbonneau, S.; Fazal, A.; Weihofen, W.A.; Gleim, S.; Salcius, M.; Sigoillot, F.; Henault, M.; Carl, S.H.; et al. CPSF3-dependent pre-mRNA processing as a druggable node in AML and Ewing’s sarcoma. Nat. Chem. Biol. 2020, 16, 50–59. [Google Scholar] [CrossRef] [PubMed]
  296. Mackey, J.R.; Lai, J.; Chauhan, U.; Beauchamp, E.; Dong, W.-F.; Glubrecht, D.; Sim, Y.-W.; Ghosh, S.; Bigras, G.; Lai, R.; et al. N-myristoyltransferase proteins in breast cancer: Prognostic relevance and validation as a new drug target. Breast Cancer Res. Treat. 2021, 186, 79–87. [Google Scholar] [CrossRef] [PubMed]
  297. Varikuti, S.; Jha, B.K.; Volpedo, G.; Ryan, N.M.; Halsey, G.; Hamza, O.M.; McGwire, B.S.; Satoskar, A.R. Host-Directed Drug Therapies for Neglected Tropical Diseases Caused by Protozoan Parasites. Front. Microbiol. 2018, 9, 2655. [Google Scholar] [CrossRef] [PubMed]
  298. Thacker, S.G.; McWilliams, I.L.; Bonnet, B.; Halie, L.; Beaucage, S.; Rachuri, S.; Dey, R.; Duncan, R.; Modabber, F.; Robinson, S.; et al. CpG ODN D35 improves the response to abbreviated low-dose pentavalent antimonial treatment in non-human primate model of cutaneous leishmaniasis. PLoS Negl. Trop. Dis. 2020, 14, e0008050. [Google Scholar] [CrossRef] [Green Version]
  299. Arnold, C. Inside the nascent industry of AI-designed drugs. Nat. Med. 2023, 29, 1292–1295. [Google Scholar] [CrossRef]
Figure 1. Distribution map for human African trypanosomiasis. The map illustrates the distribution of R-HAT and G-HAT (black). The bold line is a demarcation separating the regions in which T. b. rhodesiense and T. b. gambiense occur. The red asterisk serves to highlight Uganda, the only country in which both subspecies are found [13].
Figure 1. Distribution map for human African trypanosomiasis. The map illustrates the distribution of R-HAT and G-HAT (black). The bold line is a demarcation separating the regions in which T. b. rhodesiense and T. b. gambiense occur. The red asterisk serves to highlight Uganda, the only country in which both subspecies are found [13].
Ijms 24 12529 g001
Figure 2. Transmission and lifecycle of the T. brucei parasite. The trypanosome is digenetic, shuttling between the tsetse fly and a mammalian host. (1) An uninfected tsetse fly takes up the non-proliferative short stumpy bloodstream trypomastigotes during its bloodmeal. The bloodstream trypomastigotes form procyclic trypomastigotes that multiply by binary fission. (2 and 3) The procyclic trypomastigotes transform into epimastigotes in the salivary glands where they differentiate into the human infective metacyclic trypomastigote. (4) The bite of an infected tsetse fly injects the metacyclic trypomastigote into its next mammalian host. (5) The trypomastigote is transformed into the bloodstream form, which multiplies by binary fission and spreads in the body fluids. The bloodstream forms are either present as long slender or short stumpy forms. The blue lines represent the flagellum as protruding from the flagellar pocket (shown in yellow). The nuclei of the various morphotypes are shown in red. Adapted from [4].
Figure 2. Transmission and lifecycle of the T. brucei parasite. The trypanosome is digenetic, shuttling between the tsetse fly and a mammalian host. (1) An uninfected tsetse fly takes up the non-proliferative short stumpy bloodstream trypomastigotes during its bloodmeal. The bloodstream trypomastigotes form procyclic trypomastigotes that multiply by binary fission. (2 and 3) The procyclic trypomastigotes transform into epimastigotes in the salivary glands where they differentiate into the human infective metacyclic trypomastigote. (4) The bite of an infected tsetse fly injects the metacyclic trypomastigote into its next mammalian host. (5) The trypomastigote is transformed into the bloodstream form, which multiplies by binary fission and spreads in the body fluids. The bloodstream forms are either present as long slender or short stumpy forms. The blue lines represent the flagellum as protruding from the flagellar pocket (shown in yellow). The nuclei of the various morphotypes are shown in red. Adapted from [4].
Ijms 24 12529 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jamabo, M.; Mahlalela, M.; Edkins, A.L.; Boshoff, A. Tackling Sleeping Sickness: Current and Promising Therapeutics and Treatment Strategies. Int. J. Mol. Sci. 2023, 24, 12529. https://doi.org/10.3390/ijms241512529

AMA Style

Jamabo M, Mahlalela M, Edkins AL, Boshoff A. Tackling Sleeping Sickness: Current and Promising Therapeutics and Treatment Strategies. International Journal of Molecular Sciences. 2023; 24(15):12529. https://doi.org/10.3390/ijms241512529

Chicago/Turabian Style

Jamabo, Miebaka, Maduma Mahlalela, Adrienne L. Edkins, and Aileen Boshoff. 2023. "Tackling Sleeping Sickness: Current and Promising Therapeutics and Treatment Strategies" International Journal of Molecular Sciences 24, no. 15: 12529. https://doi.org/10.3390/ijms241512529

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