Skip to main content
Advertisement
  • Loading metrics

The Trypanosoma brucei MISP family of invariant proteins is co-expressed with BARP as triple helical bundle structures on the surface of salivary gland forms, but is dispensable for parasite development within the tsetse vector

  • Aitor Casas-Sanchez,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Software, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliations Department of Vector Biology, Liverpool School of Tropical Medicine, Liverpool, United Kingdom, Department of Tropical Disease Biology, Liverpool School of Tropical Medicine, Liverpool, United Kingdom

  • Raghavendran Ramaswamy ,

    Contributed equally to this work with: Raghavendran Ramaswamy, Samïrah Perally

    Roles Data curation, Formal analysis, Investigation, Validation, Visualization, Writing – review & editing

    Affiliation Department of Biochemistry and Microbiology, University of Victoria, Victoria, Canada

  • Samïrah Perally ,

    Contributed equally to this work with: Raghavendran Ramaswamy, Samïrah Perally

    Roles Data curation, Formal analysis, Investigation, Methodology, Writing – review & editing

    Affiliation Department of Tropical Disease Biology, Liverpool School of Tropical Medicine, Liverpool, United Kingdom

  • Lee R. Haines,

    Roles Conceptualization, Data curation, Investigation, Methodology, Validation, Visualization, Writing – review & editing

    Current address: Department of Biological Sciences, Notre Dame University, South Bend, Indiana, United States of America

    Affiliation Department of Vector Biology, Liverpool School of Tropical Medicine, Liverpool, United Kingdom

  • Clair Rose,

    Roles Formal analysis, Investigation, Methodology, Validation, Visualization

    Affiliation Department of Vector Biology, Liverpool School of Tropical Medicine, Liverpool, United Kingdom

  • Marcela Aguilera-Flores,

    Roles Data curation, Methodology, Validation

    Affiliation Border Biomedical Research Center, Department of Biological Sciences, University of Texas at El Paso, El Paso, Texas, United States of America

  • Susana Portillo,

    Roles Investigation, Methodology, Validation

    Current address: Center for Vaccine Development and Global Health, University of Maryland School of Medicine, Baltimore, Maryland, United States of America

    Affiliation Border Biomedical Research Center, Department of Biological Sciences, University of Texas at El Paso, El Paso, Texas, United States of America

  • Margot Verbeelen,

    Roles Investigation, Methodology, Validation

    Affiliation Institute of Tropical Medicine, Antwerp, Belgium

  • Shahid Hussain,

    Roles Investigation, Methodology, Validation

    Affiliation Institute of Tropical Medicine, Antwerp, Belgium

  • Laura Smithson,

    Roles Formal analysis, Investigation, Methodology, Validation

    Current address: Department of Molecular Microbiology and Immunology, Brown University, Providence, Rhode Island, United States of America

    Affiliation Biological and Medical Sciences, Oxford Brookes University, Oxford, United Kingdom

  • Cristina Yunta,

    Roles Data curation, Formal analysis, Validation

    Affiliation Department of Vector Biology, Liverpool School of Tropical Medicine, Liverpool, United Kingdom

  • Michael J. Lehane,

    Roles Formal analysis, Resources, Supervision

    Affiliation Department of Vector Biology, Liverpool School of Tropical Medicine, Liverpool, United Kingdom

  • Sue Vaughan,

    Roles Investigation, Methodology, Supervision

    Affiliation Biological and Medical Sciences, Oxford Brookes University, Oxford, United Kingdom

  • Jan van den Abbeele,

    Roles Investigation, Methodology, Resources, Supervision, Validation, Visualization

    Affiliation Institute of Tropical Medicine, Antwerp, Belgium

  • Igor C. Almeida,

    Roles Data curation, Formal analysis, Funding acquisition, Software, Supervision, Validation

    Affiliation Border Biomedical Research Center, Department of Biological Sciences, University of Texas at El Paso, El Paso, Texas, United States of America

  • Martin J. Boulanger ,

    Roles Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Supervision, Validation, Visualization, Writing – review & editing

    mboulang@uvic.ca (MJB); aacosta3@nd.edu (AAS)

    Affiliation Department of Biochemistry and Microbiology, University of Victoria, Victoria, Canada

  •  [ ... ],
  • Álvaro Acosta-Serrano

    Roles Conceptualization, Data curation, Funding acquisition, Project administration, Resources, Supervision, Validation, Writing – original draft, Writing – review & editing

    mboulang@uvic.ca (MJB); aacosta3@nd.edu (AAS)

    Current address: Department of Biological Sciences, Notre Dame University, South Bend, Indiana, United States of America

    Affiliations Department of Vector Biology, Liverpool School of Tropical Medicine, Liverpool, United Kingdom, Department of Tropical Disease Biology, Liverpool School of Tropical Medicine, Liverpool, United Kingdom

  • [ view all ]
  • [ view less ]

Abstract

Trypanosoma brucei spp. develop into mammalian-infectious metacyclic trypomastigotes inside tsetse salivary glands. Besides acquiring a variant surface glycoprotein (VSG) coat, little is known about the metacyclic expression of invariant surface antigens. Proteomic analyses of saliva from T. brucei-infected tsetse flies identified, in addition to VSG and Brucei Alanine-Rich Protein (BARP) peptides, a family of glycosylphosphatidylinositol (GPI)-anchored surface proteins herein named as Metacyclic Invariant Surface Proteins (MISP) because of its predominant expression on the surface of metacyclic trypomastigotes. The MISP family is encoded by five paralog genes with >80% protein identity, which are exclusively expressed by salivary gland stages of the parasite and peak in metacyclic stage, as shown by confocal microscopy and immuno-high resolution scanning electron microscopy. Crystallographic analysis of a MISP isoform (MISP360) and a high confidence model of BARP revealed a triple helical bundle architecture commonly found in other trypanosome surface proteins. Molecular modelling combined with live fluorescent microscopy suggests that MISP N-termini are potentially extended above the metacyclic VSG coat, and thus could be tested as a transmission-blocking vaccine target. However, vaccination with recombinant MISP360 isoform did not protect mice against a T. brucei infectious tsetse bite. Lastly, both CRISPR-Cas9-driven knock out and RNAi knock down of all MISP paralogues suggest they are not essential for parasite development in the tsetse vector. We suggest MISP may be relevant during trypanosome transmission or establishment in the vertebrate’s skin.

Author summary

The Trypanosoma brucei group of parasites are exclusively transmitted to the vertebrate host by the tsetse vector alongside insect saliva. To better understand trypanosome transmission, we investigated the protein composition of T. brucei-infected tsetse saliva using a mass spectrometry proteomics approach. We found that, in addition to proteins from tsetse saliva and Sodalis glossinidius (a bacterial tsetse symbiont), trypanosome-infected saliva contains several parasite surface glycoproteins, including a partially characterized family of invariant proteins herein named Metacyclic Invariant Surface Proteins (MISP). We show that MISP is primarily expressed, together with metacyclic Variant Surface Glycoprotein (mVSG) and Brucei Acidic Repetitive Protein (BARP), on the surface of the infectious metacyclic stage of T. brucei. Its triple helix bundle architecture appears tethered to the outer membrane by an extended glycosylphosphatidylinositol-anchored C-terminal tail that putatively projects MISP above the VSG coat. Our findings provide new insights into the surface architecture of the T. brucei metacyclic stage and describe the challenges associated with developing transmission-blocking vaccines against tsetse-transmitted trypanosomes.

Introduction

African trypanosomes are the causative agents of human African trypanosomiasis (HAT) or sleeping sickness and animal African trypanosomiasis (AAT or Nagana disease) in most sub-Saharan countries. Multiple trypanosome species, including the group of Trypanosoma brucei parasites (i.e. T. b. gambiense, T. b. rhodesiense, and T. b. brucei), are transmitted by several tsetse species (Glossina spp.). Despite decreasing HAT cases since 2010, due largely to a program of active screening and treatment of the human population together with an efficient vector control program, AAT continues to cause significant economic losses in the agricultural sector with >1 million cattle dying each year [1]. Since these parasites undergo antigenic variation to effectively evade the host’s immune system, preventive vaccines have been difficult to develop, and only one has recently shown protection against T. vivax [2].

During its life cycle, T. brucei undergoes many developmental changes to adapt to the different environments within the mammalian host and the tsetse vector [3]. All T. brucei life stages have in common the expression of major surface glycosylphosphatidylinositol (GPI)-anchored glycoproteins known to be important for parasite survival. In bloodstream form (BSF) stages, the variant surface glycoprotein (VSG) coat is responsible for antigenic variation [46], clearance of host immunoglobulins bound to VSGs [7] and preventing antibodies from binding to conserved VSG epitopes and invariant surface proteins (ISG), such as transporters and receptors [811]. When ‘stumpy’ BSF trypanosomes are ingested by tsetse, they quickly (24 to 48 hours) differentiate into the procyclic trypomastigote stage within the fly’s midgut [3,12]. During differentiation, VSGs are no longer expressed and the parasites lose the VSG coat via the concerted action of major surface metalloproteases and the GPI-phospholipase C (GPI-PLC) [13,14]—a process that appears to modulate the tsetse’s immunity in favor of parasite infection [15]. Concomitant with VSG disappearance, the parasites express a new coat comprised of GPEET- and EP-procyclin isoforms once the infection establishes in the midgut [1618] and proventriculus (PV) [19]. Within the tsetse midgut, procyclic trypomastigotes first colonize the ectoperitrophic space of the anterior midgut, from which they further migrate to the tsetse cardia or PV [20]; alternatively, both the anterior midgut and PV can be simultaneously colonized as trypanosomes reach the ectoperitrophic space after crossing an immature peritrophic matrix at the PV [19]. After 10–15 days post-infection, depending on parasite strain, trypanosomes differentiate into epimastigote forms (EMFs) within the PV, which then migrate to the salivary glands (SGs), attach to the epithelial cell microvilli and switch on the expression of Brucei Alanine-Rich Proteins (BARP) [21]. The attached EMFs further differentiate into pre-metacyclic trypomastigotes [22,23], during which time BARP expression is lost, a new coat of metacyclic VSG (mVSG) develops [24,25], and cells detach from the SG epithelium. Finally, mature metacyclic trypomastigote forms (MCFs) [23,26] are inoculated along with saliva into the skin of a vertebrate host in a subsequent feed. Unlike in BSFs, the surface composition of T. brucei MCFs is less well characterized primarily due to the lack of an efficient in vitro culture system of this life stage, the low percentage of flies with salivary gland infections, and the low yields of MCFs from infected salivary glands.

In contrast to BSFs that possess a large repertoire of vsg genes [27,28], MCFs express a single mVSG protein (also called metacyclic variant antigen type; mVAT) from a much-reduced gene repertoire. Based on antibody recognition, it is estimated that a population of tsetse-derived MCFs express different mVSGs (as many as 27) to ensure infection of hosts that may have been pre-exposed to any of these antigenic types [29,30]. Given that MCFs develop in the tsetse SGs and their persistence in the host after transmission is relatively short [26], they do not seem pressured to undergo antigenic variation. Despite the differences in the expression mechanisms between VSGs and mVSGs, it is assumed that their function and structural organization is equivalent. BSF VSGs form a protective outer layer that covers the parasite surface (~5 million homodimers/cell) [31] to mask the invariant surface proteins from the host immune system. Many of these invariant proteins are predicted to be shorter in height than the VSG homodimers and to be attached to the plasma membrane via transmembrane domains [8,11]. The exceptions to this are the haptoglobin-hemoglobin (HpHbR) and transferrin receptors, which are GPI-anchored and exclusively localize in the flagellar pocket [9,10].

During a blood meal, the infected tsetse inoculates MCFs into the host along with saliva [26,32] to inhibit the host hemostatic response to the cutaneous trauma [33,34]. However, during parasite development in the tsetse SGs, there is a drastic (~80%) transcriptional down regulation of genes encoding for salivary proteins, which induces a feeding phenotype that may amplify disease transmission [32,35,36]. Although the proteome of tsetse saliva from T. brucei-infected flies has been determined [34], soluble parasite factors have not been characterized in detail and thus, the molecular mechanism by which these parasites manipulate the vector to promote transmission is largely unknown. Here, we identified trypanosome proteins from infected tsetse saliva using a high-resolution mass spectrometry-based proteomics approach. We found that T. brucei-infected saliva is particularly enriched with peptides from trypanosome surface proteins such as BARP, mVSG and a partially characterized family of glycoproteins herein named Metacyclic Invariant Surface Proteins (MISP). We show evidence that MISP, previously named as salivary gland epimastigote 1 (SGE1) [37, 38], are a small family of invariant glycoproteins primarily expressed on the surface of the metacyclic stage of T. brucei. Structural data reveal a triple helix bundle architecture for MISP that appears to be tethered to the outer membrane by an extended, unstructured C-terminal tail, potentially projecting MISP above the mVSG coat and thereby allowing antibody recognition. Our findings provide a new insight into the surface architecture of MCFs and are discussed with respect to the molecular crosstalk between parasite and vector, and the challenges associated with developing transmission-blocking vaccines against tsetse-transmitted trypanosomes.

Results

Proteomic analysis of T. brucei-infected tsetse saliva reveals an enrichment of parasite GPI-anchored surface proteins

We initially investigated the protein composition of Glossina morsitans morsitans saliva with a T. b. brucei infection in the salivary glands to identify novel soluble factors important for parasite transmission. We performed high-resolution nano-liquid chromatography-tandem mass spectrometry (nLC-MS/MS) on pools of saliva from flies at 30 days post-infection (dpi) in three biological replicates (S1 Fig). After filtering out the Glossina protein hits (S1 Table), we detected peptides from the bacterial symbiont Sodalis glossinidius [39] in both uninfected and trypanosome-infected saliva (S2 Table). The presence of these symbiont peptides was validated by immunoblotting using an anti-Sodalis Hsp60 monoclonal antibody (S2 Fig). Following manual curation of the MS/MS spectra, we also identified 45 unique peptides derived from a total of 27 T. b. brucei proteins in infected saliva datasets only (>95% confidence protein identification) (S3 Table). Of the identified proteins, 17 (62.9%) were predicted to be intracellular (S4 Table) mainly linked to functions such as protein folding (14.8%), cytoskeleton (11.1%), and metabolic processes (7.4%). The remaining 10 proteins (37.1%) were predicted to be GPI-anchored surface proteins from three major families (Fig 1A). We found four unique peptides shared by eight different BARP isoforms and identified five different mVSGs from seven unique peptides (Fig 1A, S1 File and S3 Table). Notably, we also identified a single peptide (100% confidence, supported by two spectra, S1 File and S3 Table) belonging to SGE1 [37], a small family of GPI-anchored proteins we characterized and renamed metacyclic invariant surface proteins (MISP) (Fig 1A) because of its predominant life-cycle expression (see below). The conserved peptide ‘SVAEDNSAASTAR’ was fully conserved across all MISP isoforms (Fig 1B–1D and S3 Table).

thumbnail
Fig 1. Identification of sequence analyses of MISP.

(A) Summary table of trypanosome GPI-anchored proteins identified in T. brucei-infected tsetse saliva by nLC-MS/MS. All identified proteins and peptides have >95% confidence and were manually curated. Protein accession code (Protein ID), protein description (Annotation), number of identified unique peptides (Peptides), and number of spectra (Spectra) are indicated; see peptides and spectra in S1 File. Accession codes either from TriTrypDB (BARP and MISP) or UniProt (VSG). (I) representative ID for Tb927.9.15520, Tb927.9.15530, Tb927.9.15550, Tb927.9.15590, Tb927.9.15600, Tb927.9.15620 and Tb927.9.15660; (II) representative ID for VSGs 646, 769, 3613, 1125.408, 1125.474, 1125.1142, 1125.4207 and 1125.4707; (III) representative ID for Tb927.7.360, Tb927.7.380, Tb927.7.400, Tb927.7.420, Tb927.7.440. (B) Phylogenetic tree generated from a multiple sequence alignment of all MISP predicted proteins. Note the demarcation of two sub-families MISP-A (light blue) and MISP-B (dark blue) and the least-conserved T. congolense isoform (red). The T. b. brucei MISP are boxed. Tree created using PhyML 3.1 and the WAG substitution model of maximum likelihood; bootstrap of 500; numbers show branch support values (%); scale bar: 0.1 substitutions/site. (C) Schematic representation of the T. b. brucei chromosome 7 region containing the misp gene array. Genes represented by arrows; misp-A (light blue), misp-B (light purple), unrelated genes (light grey); C-terminal repeats represented by purple rectangles on the arrows. (D) Multiple protein alignment of T. b. brucei MISP isoforms. Residues colored from grey (full conservation) to bright red (highest divergence). Protein domains highlighted: N- signal peptide (green), predicted N-glycosylation sites (yellow asparagine), disulphide bonds identified in crystal structure (cysteines in black boxes), C-terminus repetitive motives of 26 residues (blue) and GPI-anchor signal peptide (red). The conserved peptide ‘SVAEDNSAASTAR’ identified by nLC-MS/MS is highlighted in orange. The top bar numbers indicate residue position starting from Met1. Top bar colors indicate the main protein domains N-terminus (purple) and C-terminus (grey).

https://doi.org/10.1371/journal.ppat.1011269.g001

The Metacyclic Invariant Surface Proteins (MISP) family

MISP proteins are part of the large Fam50 family of trypanosome surface proteins [40]. T. b. brucei contains five misp paralog genes (Tb927.7.360, Tb927.7.380, Tb927.7.400, Tb927.7.420, Tb927.7.440 in reference strain TREU927), which encode for proteins that are unique in sequence and have no predicted function. All five T. b. brucei paralogs are localized to chromosome 7 (Fig 1C) and are separated by unrelated genes encoding hypothetical proteins (Tb927.7.370, Tb927.7.390, Tb927.7.410 and Tb927.7.430). Notably, the same misp homologs are found in T. b. gambiense (Tbg972.7.270 and Tbg972.7.290). Furthermore, slightly divergent versions of misp genes are found in the animal trypanosome species T. evansi (TevSTIB805.7.300, TevSTIB805.7.320 and TevSTIB805.7.380) and T. congolense (TcIL3000.0.02370), and no homologs were found in the T. vivax genome (S3 Fig). Using a phylogenetic analysis (PhyML 3.1) [41], we determined that MISP are comprised of two sub-families (MISP-A and MISP-B) (Fig 1B). MISP-A members are mainly characterized by containing one to three 26 residue repetitive motives at the C-terminus (see below) and MISP-B isoforms, in contrast, present no C-terminal repeats. The T. congolense homolog (TcIL3000.0.02370) shares only 34.2% identity with the others and was not included into either sub-family (S3B Fig). A comparison of all T. b. brucei MISP paralogs revealed an overall amino acid sequence identity of ~79% (Fig 1D), with the N-terminal domain (Met1–Glu237) showing greater conservation (88.7% identity) compared to the divergent C-terminal (Lys238-Ser357) (49.0% identity). Importantly, sequence divergence is not random but primarily localized to the charged 26 residue repetitive motives (EVEEVPKKDPEGNVEVPEDKEKTERT or DVQEISREEFEGNVEVPEDKEKTERT). We identified predicted signal peptides for all MISP isoforms (Met1–Ala/Thr17). While no transmembrane domains were identified, all MISP isoforms contain a predicted GPI omega site at the C-terminus (Fig 1D and S5 Table). Furthermore, all MISP isoforms contain at least one potential N-glycosylation site (Asn59), while Tb927.7.400 and Tb927.7.420 contain an additional site on Asn121 (S5 Table). Overall, sequence-based predictions suggest that MISP is a family of surface GPI-anchored glycoproteins presenting little sequence divergence and differential C-terminal lengths.

MISP are expressed on the surface of T. b. brucei salivary gland stages

To understand the expression pattern of individual misp paralogs throughout the T. b. brucei’s life cycle, we used semi-quantitative RT-PCR to detect transcript levels of individual misp isoforms and barp, which are markers for salivary gland epimastigotes [21] (Fig 2A). Total RNA was purified from all isolated T. b. brucei life stages obtained from both in vitro cultures (i.e., bloodstream forms and procyclic cultured forms) and tsetse-derived parasites at 30 dpi. (i.e., midgut procyclics, proventricular parasites (mixture of 95% mesocyclics and 5% epimastigotes), whole infected SGs, and metacyclics isolated from infected saliva (mixture of 95% metacyclics and 5% epimastigotes). Thirty day old uninfected flies were used as an age-matched control. Due to the high sequence identity among misp paralogs, assessing the expression of individual misp transcripts by real-time RT-PCR was not feasible and hence, we used semi-quantitative RT-PCR to target the different C-terminal misp repetitive regions, which are unique amongst four of the paralogs with differences of 78 bp (S4 Fig). While all control samples from uninfected flies were negative, whole infected SGs and purified metacyclics from saliva showed the highest levels of misp transcripts compared to midgut procyclics (11.8-fold) and BSFs (25.1-fold) (Fig 2A). Expression of misp transcripts was also found to be high in trypanosomes colonizing the PV, suggesting it is in those stages where gene up regulation starts. Although there seems to be an overall preferential selection for paralogs with the shortest C-termini (misp400/420), the longer paralogs misp380 and misp360 were up regulated and significantly higher in SG stages. In parallel, the transcriptional profile of misp was validated by measuring that from a transgenic cell line overexpressing misp360 under a tetracycline-inducible promoter (S4 Fig). Interestingly, all barp transcripts (originated from 14 paralogs sharing >60% identity, S5 Fig) followed a similar expression pattern, although PV-associated trypanosomes were the highest expressors, suggesting that the expression of barp may precede that of misp genes during development.

thumbnail
Fig 2. Expression and localization of MISP and BARP.

(A) Relative mRNA expression levels (arbitrary units, normalized to the expression of tert) of T. b. brucei misp paralogs (blue) and barp (yellow), in cultured bloodstream forms (BSF), procyclic cultured forms (PCF), midgut procyclics (MG), proventricular forms (PV), salivary gland forms (SG) and purified metacyclic forms (MCF). Bars represent mean values; error bars represent ±S.D from two technical replicates. (B) Representative images of non-permeabilized fixed parasites extracted from infected tsetse salivary glands, immunostained with either anti-MISP (B) or anti-BARP (C) polyclonal antibodies; epimastigote (EMF), pre-metacyclic (P-MCF) and metacyclic forms (MCF); shown as phase, anti-MISP (cyan) or anti-BARP (yellow), and DAPI (magenta); kinetoplasts (K) and nuclei (N) are annotated in the DAPI channel; anterior (A) and posterior (P) ends of the parasite are annotated in the Phase channel. (D) Maximum intensity projection of representative metacyclic cells immunostained with anti-MISP (top) or anti-BARP (bottom) polyclonal antibodies and 3D-reconstructed from z-stacks; y-z orthogonal views show surface localization of the signal and graphs on the right plot fluorescence intensity over the white line crossing the cell section. (E) Localization of anti-HA monoclonal antibodies on metacyclic (MCF), bloodstream form (BSF) and procyclic cultured form (PCF) transgenic T. b. brucei cells overexpressing ectopic HA-tagged MISP360 upon doxycycline induction; shown as phase, anti-HA (cyan), DAPI (magenta) and merge; kinetoplasts (K) and nuclei (N) are annotated in the DAPI channel; anterior (A) and posterior (P) ends of the parasite are annotated in the Phase channel. (F) Representative HRSEM images of wild type metacyclic (MCF), procyclic MISP360 overexpressor uninduced (PCF MISP -) and induced (PCF MISP+) cell surfaces immunostained with anti-MISP polyclonal antibodies; secondary electron detection (SE), 50% inverted backscatter electron detection (50% BSD); scale bars annotated in SE.

https://doi.org/10.1371/journal.ppat.1011269.g002

Having identified SG stages as the major expressors of misp and barp transcripts, we next determined which life stages translated these transcripts into proteins by immunostaining with polyclonal antibodies that recognize epitopes conserved across all MISP or BARP isoforms. Antibody binding was only to SG parasites, both within the glands (S6 Fig) and ex vivo (Fig 2B and 2C), and no recognition was detected on BSFs, PCFs, midgut procyclics or PV parasites by either antibody (S7 Fig). Surprisingly, despite PV-associated trypanosomes showing high levels of misp and barp transcripts, the corresponding gene products are not produced until they reach the next developmental stage in the salivary glands. Instead, we confirmed that both mesocyclic and epimastigote forms in the proventriculus expressed EP-procyclin as the major surface GPI-anchored protein (S8 Fig). Furthermore, both polyclonal anti-MISP and anti-BARP antibodies recognized all parasite cells and stages in infected SGs, including attached epimastigotes, pre-metacyclics and mature metacyclic forms (Figs 2B, 2D and S6). The relative mean fluorescence intensity of MISP was higher in metacyclics compared to epimastigotes (2.3-fold); in contrast, BARP intensity was found to be higher in epimastigotes than in metacyclics (3.3-fold) (S9 Fig). Moreover, MISP appeared to localize evenly across the cell surface of non-permeabilized fixed epimastigote and metacyclic cells (Fig 2B). In contrast, BARP appears homogeneously expressed on the surface of all epimastigotes, yet on metacyclics, BARP is detected on the surface of few (~10.4% of the cells; n = 481), while the rest show exclusively flagellar expression (Figs 2C and S10). To confirm the identity of SG-specific trypanosomes, we measured the length of parasites and the relative position of kDNA and nucleus (S11 Fig), and performed immunostainings with anti-CRD polyclonal antibodies, which recognize the GPI cross-reacting determinant (CRD) epitope [23,42]. While epimastigotes and pre-metacyclics were not detected by the antibody, mature metacyclics showed surface staining as indirect detection of mVSGs that are partially cleaved by GPI-PLC (S12 Fig). To confirm the localization of MISP and BARP in metacyclics, immuno-stained cells were 3D-rendered using confocal laser scanning microscopy and orthogonal views from z-stacks confirmed the fluorescence only localized to the edges of the cell body (Fig 2D).

As an additional approach to investigate MISP localization, we engineered a T. b. brucei transgenic cell line to ectopically express HA/eGFP-tagged MISP360 under a tetracycline-inducible promoter (Figs 2E and S13). Upon induction, ectopic MISP localized on the metacyclic cell body surface in a pattern resembling that of wild type cells (S13C Fig). However, in BSF, ectopic MISP was only detected in the flagellar pocket and, upon differentiation to PCF, it was then found redistributed to the flagellum and co-localised with the paraflagellar rod (PFR) (S13C Fig). We also used high-resolution scanning electron microscopy (HRSEM) to precisely locate anti-MISP antibodies on the surface of stained cells (Fig 2F). While uninduced MISP360 overexpressing PCF cells did not show antibody binding, induction of the transgene led to detection of the ectopic MISP mainly along the flagellum as in Fig 2E. In contrast, native MISP was equally detected in wild type metacyclic cells on both flagellum and cell body, thus validating the observations from confocal microscopy (Fig 2B and 2D). To confirm exposure of MISP epitopes in native conditions and free of potential fixative-derived artifacts, we performed immunostaining on live wild type metacyclic cells, which still displayed positive binding of anti-MISP antibodies on the cell surface (S14 Fig). Lastly, metacyclics extracted from flies infected with T. brucei TSW196 strain are also recognized by the anti-MISP antibody (S15 Fig), confirming that staining is not specific for AnTat 1.1 metacyclic trypanosomes.

The N-terminal ectodomains of MISP and BARP adopt a triple helix bundle structure

To complement the expression and localization analyses, we next determined the crystal structure of the conserved N-terminal ectodomain (Gly24-Ala234) of MISP360 (Tb427.07.360; T. b. brucei MISP360 homolog in reference strain Lister 427) to 1.82 Å resolution (Fig 3). Crystals of purified, monomeric MISP360 (Figs 3A–3B and S16A–S16B) were obtained using the hanging drop method and grew in space group P212121 with a single molecule in the asymmetric unit. The structure of MISP360 was determined by molecular replacement using a truncated form of T. congolense GARP (PDB entry: 2Y44) as the search model. The identification of GARP as a suitable model was based on secondary structure predictions as the amino acid sequence identity of the N-terminal ectodomain is only 15%. The overall structure of MISP360 was well defined with only two residues from the N-terminus remaining unmodelled (Gly24 and Ser25). The core of the N-terminal ectodomain adopts an elongated structure measuring approximately 83 Å in height and spanning approximately 26 Å in width (Fig 3C, left). The ectodomain is well ordered with low B-factors throughout the protein (Fig 3C, middle). MISP360 adopts an overall triple helical bundle structure composed of a core of extended twisted helices capped by a smaller helical bundle at the N-terminal end. The helical bundle that dominates the structure consists principally of three helices (Fig 3C, right): helix I is comprised of residues Val29-Ser83, helix II of Glu88-Thr130 and helix VI of Phe182-Ala233. The three helices adopt a bend of approximately 30 degrees at Gly44 (helix I), Ala127 (helix II) and Leu195 (helix VI) and collectively form the helical bundle cap. In addition to the terminal portions of the three major helices, the helical bundle includes three shorter helices: helix III (Asp134-Glu142), which is connected by a 5-residue loop to helix IV (Ser148-Gly156), and helix V (Ser166-Phe179) connected to helix IV by a 9-residue loop. Helices IV and V lie at either side of helix I and form the broadest face of the helical bundle cap. Moreover, the head structure of MISP360 is anchored by two disulphide bonds; one between Cys36 and Cys157, and the other between Cys177 and Cys185 (Fig 1D).

thumbnail
Fig 3. Crystal structures of MISP and BARP N-terminal ectodomains.

(A) Schematic representation of MISP360 and BARP protein sequences, highlighting in color the region recombinantly expressed in E. coli and crystallized; numbers indicate residue positions, signal peptide (SP), C-terminal domain (C-term), GPI-anchor peptide (GPI). (B) FPLC SEC elution chromatogram of recombinant MISP360; SDS-PAGE of monomeric recombinant MISP360 (inset). (C) MISP360 N-terminus crystal structure with molecular surface representation. The structure was found to be 83.3 Å tall, 25.5 Å in width with a 30° helix bend at the top (left); B-factor putty model with ordered regions displayed in blue and thin tubes and flexible regions in red and thicker tubes (middle); secondary structure depiction highlighting the organization of the helices (right); bend-forming residues are in parentheses; coiled structures are represented in white. (D) BARP high confidence homology model generated with Modeller 9v18 [43] shown with a semi-transparent surface representation. The BARP model reveals an overall architecture that is 82.7 Å tall, 25.0 Å in width with a 34° helix bend consistent with the molecular envelope calculated from the 3.2 Å X-ray crystallography diffraction data set. (E) Structural superimposition of the MISP360 N-terminus crystal structure (cyan) with GARP (purple, PDB: 2Y44) and the high confidence model of BARP (orange) (left). A black dashed line highlights the area where the apical surface pocket is found. Top view of an electrostatic surface representation of MISP360, BARP and GARP where the surface pocket is found (middle), and its orthogonal view indicating the residues forming the pocket and size (right). (F) Conservation of the pocket-forming (top) and overall (bottom) residues across MISP360, BARP and GARP depicted on the MISP360 structure (left). Residue conservation between the five T. brucei MISP isoforms on the MISP360 structure (right); from most variable (dark blue) to most conserved (dark purple). (G) Schematic comparison of the C-terminal domain lengths of BARP, VSG and MISP; scale bar = 10 residues; dashed line sets length baseline (VSG). (H) Structural model of the proposed T. brucei metacyclic surface glycocalyx displaying the GPI-anchored metacyclic VSG homodimers (mVSG; light and deep purple), MISP-A (cyan), MISP-B (dark blue) and remains of BARP (yellow). The mVSG structure is represented with a model of mVAT4 based on the crystal structure of VSG 221 N- (PDB: 1VSG) and C-terminus (PDB: 1XU6). MISP is represented with the crystal structure of MISP360 N-terminus (PDB: 5VTL) and a model of its C-terminus; BARP is represented with our confidence model.

https://doi.org/10.1371/journal.ppat.1011269.g003

Given that BARP is co-expressed with MISP in both epimastigote and metacyclic forms (Fig 2C), we also sought to determine its crystal structure after expression in E. coli and purification (S16C and S16D Fig). Diffraction data was collected to 3.2 Å resolution and structure solution revealed a molecular envelope consistent with MISP. However, the data quality did not support a detailed atomic level refinement and thus we opted to generate a high-confidence BARP (Tb927.9.15630) homology model using Modeller 9v18 [43]. Consistent with the X-ray data, the BARP model revealed an extended helical architecture with comparable dimensions and a largely conserved helical bundle relative to MISP360 (Fig 3D).

The lack of significant sequence identity between MISP360 and any protein with known function led us to perform a DALI search [44] to identify structural homologs. T. congolense’s GARP, of unknown function, was identified as the top hit with a Z-score of 21.4. A least squares superposition between the two structures resulted in a root-mean-square deviation of 1.7 Å over 183 Cα atoms (Fig 3E, left). The DALI search also revealed structural homology to the previously characterized haptoglobin-hemoglobin receptor (HpHbR) from T. brucei (PDB: 4X0J) and T. congolense (PDB: 4E40) [45, 46], and to a T. brucei VSG monomer (PDB: 1VSG) with Z-scores of 18.1, 17.7 and 6.5, respectively. All these proteins exhibit a complementary core of twisted three helical bundles (S17 Fig). Structural comparison, however, indicates a closer architectural similarity with HpHbR compared to VSG, due in part to the breakdown of the third helix into loops and extensions enabling substantial conformational diversity in VSG [47]. Moreover, MISP360 is monomeric in contrast to dimeric VSGs [47]. Despite the general structural similarity with GARP, BARP, HpHbR and the VSG monomer, the overall low sequence identity and the lack of key conserved residues suggest a different biological role for MISP.

Surface analysis of the MISP360 and GARP [48] crystal structures, and the BARP high confidence homology model revealed a similar distribution of acidic and basic patches along the entire length of the structure with no clear localized charge densities that would indicate a molecular recognition site (Fig 3E). However, a surface pocket of similar dimensions was identified at the membrane distal end near the region where the core helices bend in each protein (Fig 3E), but intriguingly, the residues forming the pockets are quite divergent. While the pocket in MISP360 is mainly formed by hydrophobic residues, BARPs contain both hydrophobic and basic residues, and GARPs are formed by acidic, polar and hydrophobic amino acids. This variability in pocket residue composition prompted us to map the residues conserved between MISP360, BARP and GARP onto the MISP360 crystal structure using ConSurf [49, 50] (Fig 3F, left), which showed no clear conservation among the pocket-forming residues. Moreover, mapping conservation between the five T. brucei MISP isoforms on MISP360, revealed that most of the conserved residues are found forming the core of the helices and pocket. However, a striking 51% of the non-conserved residues (representing 22.3% of the full protein sequence) are in close proximity to the pocket (Fig 3F, right), suggesting these residues may mediate differential affinity or recognition of putative ligands.

A structural comparison between the VSG MiTat 1.2 N-terminus (PDB: 1VSG), MISP360, and the BARP model revealed they all share a similar height (~99.9 Å, ~93.0 Å and ~89.2 Å, respectively) (S18A Fig). However, when comparing the sequence length of their unstructured C-terminal domains (Fig 3G) with that of VSGs, it was clear that MISP360 and MISP380 have ~71% and ~34% longer C-termini; MISP440 has an equivalent length and MISP400/420 have a 60% shorter C-terminus. All BARP isoforms (except Tb927.9.15590) lack a C-terminal domain. By combining the crystal structure of MISP360 and the MISP C-terminal sequences, we generated structural models for all T. brucei MISP isoforms using I-TASSER [5153] and IntFOLD [54] (S18B Fig). Similar to TcHpHbR and VSG monomers, we predict that the MISP monomeric triple helical bundle is vertically oriented and, within the MCF surface context, is in close contact with neighboring mVSG homodimers. Since MISP360/380 have longer C-terminal domains than VSGs, and despite their lack of secondary structure, we predict that the MISP N-terminal domains (except for MISP400/420) could extend above the mVSG coat (Figs 3H and S18B). Consistent with our live immunostaining results (S14 Fig), these data suggest that despite being covered by a dense coat of mVSG homodimers, metacyclic trypomastigotes still allow room for the co-exposure of invariant GPI-anchored glycoproteins throughout the cell surface like BARP and MISP.

MISP does not protect mice against a tsetse-transmitted T. brucei infection

Since we found MISP isoforms to be highly conserved, expressed on the metacyclic surface and susceptible to antibody binding in native conditions, we considered MISP as transmission-blocking vaccine candidates against T. brucei. To explore the experimental immunoprotective properties of MISP, the N-terminal ectodomain of MISP360 (Fig 4A) was recombinantly produced in E. coli, purified by nickel affinity and size exclusion chromatography (Fig 4B–4C), depleted from endotoxins, and used to formulate a vaccine with glucopyranosyl lipid adjuvant aqueous formulation (GLA-AF) as adjuvant. In parallel, we also immunized mice with recombinant BARP (Fig 4D–4F) and dead whole T. b. brucei metacyclic cells. Ten BALB/c mice per group were immunized by subcutaneous injection and received two more boost doses after 10 and 20 days (Fig 4G–4H and S2 File). Ten days after the last booster, each mouse was bitten once by a single tsetse with a confirmed SG infection. Infection was followed up to 31 days post-challenge and parasitemia was regularly measured for each individual (Fig 4G). Animals vaccinated with either MISP or BARP developed an infection and did not show a significant reduction in parasitemia compared with the control groups (PBS and adjuvant only), at any point. Only mice immunized with dead metacyclic cells showed a significant reduction of parasitemia at 7 days post-challenge (Fig 4H), but this was restored to levels comparable to the controls the day after. Despite the immunized animals successfully developing an immune response against either antigen by generating high titers of specific antibodies as shown by immunoblotting (S2 File), vaccination did not prevent the development of a T. b. brucei infection.

thumbnail
Fig 4. Immunoprotective potential of MISP and BARP.

Representation of MISP (A) and BARP (B) recombinant proteins produced for mouse vaccination; sections in gray were not included; signal peptide (SP), C-terminus domain (CTD), GPI anchor peptide (GPI); numbers indicate position from Methionine 1. Size exclusion chromatography chromatograms of recombinant MISP (C) and BARP (D), indicating the purified peaks including the SDS-PAGE of the peak fraction as inset. Western blotting of BSA as control, recombinant MISP and BARP, probed with either anti-MISP polyclonal antibodies (E) or anti-BARP (F). (G) T. brucei parasitaemia development (parasites/mL blood) in mice either vaccinated with PBS only (gray), PBS and adjuvant (dark gray), recombinant MISP and adjuvant (blue), recombinant BARP and adjuvant (yellow) or dead metacyclic cells (purple) from 5 to 32 days post-challenge; dots indicate the mean of each group (n = 10 per group). (H), parasitemia levels from 5 to 8 days post-challenge highlighting the significant (*) reduction of parasitemia at day 7 post-challenge in mice vaccinated with dead metacyclic cells.

https://doi.org/10.1371/journal.ppat.1011269.g004

MISP is not essential for trypanosome colonization of the tsetse salivary glands

To gain more insight into the essentiality of MISP, we created a T. b. brucei transgenic cell line that allows the RNAi knockdown of all misp paralogs at once through a tetracycline-inducible RNAi stem loop system [55] targeting a conserved gene region. Although misp knockdown was successfully induced, both transcript and protein downregulation were only partial and did not lead to any conclusive phenotype either in PCF in vitro or within the tsetse salivary glands (S19 Fig). We then used a CRISPR-Cas9 strategy to knock out (KO) misp by targeting DNA breaks at both 5’ and 3’ sequences conserved in all paralogs, and replacement with drug selection markers (Fig 5A). An addback (AB) rescue cell line was created by inserting a tetracycline-inducible ectopic misp360 gene for overexpression. As expected, due to the native expression pattern of misp, neither PCF cell line presented any growth or morphology phenotype in vitro. We then infected tsetse with either misp KO or AB cell lines and assessed infection rates at 30 dpi. All cell lines (i.e. KO, uninduced (AB-) and induced (AB+) addback) yielded normal midgut and proventricular infection rates, comparable to that of wild type cells. However, salivary gland infection rates of KO and AB- experienced a slight (although not statistically significant) reduction (Fig 5B) despite the observation that morphology and motility were apparently not affected, suggesting that MISP may have a minor but non-essential role in salivary gland infection establishment and/or maintenance. Lastly, misp KO cells re-expressing MISP360 (AB+ cells) failed to restore normal salivary gland infection rates. It remains to be determined whether this is due to the differential expression level compared to the endogenous gene and/or whether more than one isoform is necessary for complementation. Lack of MISP expression in KO cells was confirmed by immunostaining with anti-MISP antibody (Fig 5C), while expression of BARP and CRD remain comparable to the wild type and AB+ cells (S20 Fig).

thumbnail
Fig 5. Characterization of MISP null mutants.

(A) PCR genotyping of MISP KO clones (#1 to #6), wild type (WT), and negative template control (NTC) by amplification of MISP C-terminal domains (top gel); amplification of tert as control (bottom gel); relative molecular weight on the right (bp). (B) Infection prevalence (percentage) in tsetse salivary glands infected with either wild type (WT), MISP knock-out clone #3 (KO), MISP KO uninduced addback (AB-), and MISP KO induced addback (AB+); two biological replicates (replicate 1 (light blue circle); replicate 2 (dark blue square)); error bars indicate binomial ±SE of sample proportion (replicate 1(black), replicate 2 (gray)); sample population indicated next to each point. (C) Representative images of wild type (WT), MISP knock out (KO) and induced MISP addback (AB) immunostained with anti-MISP polyclonal antibodies (cyan), DAPI (magenta), merge, and differential interference contrast (DIC); scale bars = 5 μm.

https://doi.org/10.1371/journal.ppat.1011269.g005

Discussion

Transmission of vector-borne pathogens usually requires insect saliva as a vehicle, but it is also accompanied by a series of soluble components from both the parasite and the insect. For instance, in the sand fly, Leishmania spp. parasites secrete abundant promastigote secretory gel and exosomes that are important virulence factors for transmission and establishment of the parasite infection in the vertebrate host [5658]. Here, we performed semi-quantitative proteomics to analyze the composition of T. b. brucei-infected tsetse (Glossina m. morsitans) saliva with the aim of identifying potential soluble factors involved in parasite transmission. Alongside hits from Glossina, we found proteins from the bacterial symbiont S. glossinidius, which compared to naïve flies seem to be more abundant in saliva during a trypanosome infection, as shown by immunoblotting. This could be explained in part by a higher cell permeabilization or microbial dysbiosis led by the trypanosome infection, as Sodalis is capable of intracellular and extracellular growth within tsetse, including in the salivary glands [39]. An overall reduction in the number of salivary proteins was confirmed in infected saliva as previously observed [32,34]. In addition, we identified 27 different T. b. brucei proteins in infected saliva, of which 62.9% were cytosolic and cytoskeletal. It remains unclear why trypanosome proteins are found in tsetse saliva and how they are released by the parasite. Their source could be dying parasites, and/or a more specific mechanism of protein secretion such as shedding of extracellular vesicles (EV) [59]. Indeed, it has been demonstrated that T. brucei BSFs secrete EVs that contain several virulence factors and proteins involved in parasite persistence within the host, including serum resistance-associated protein, VSGs, adenylate cyclase (GRESAG4), and GPI-PLC [60]. Notably, we showed that T. brucei-infected saliva is particularly enriched with trypanosome GPI-anchored surface proteins (37.1%); i.e. several VSGs, BARPs, and one (Tb927.7.360) belonging to a small family of hypothetical proteins, we herein named as metacyclic invariant surface proteins (MISP). Nevertheless, it remains to be determined whether MISP, VSGs, BARP, and other proteins here found in tsetse saliva are part of EVs, secreted in soluble form, or both.

We identified and characterized the MISP family by studying gene and protein expression during parasite life cycle, cellular localization and accessibility to antibodies, protein tertiary structure, immunoprotective potential, and essentiality for parasite development in the tsetse vector. Most of these studies were conducted alongside BARP, a known family of surface proteins expressed by epimastigotes in SGs [21]. In our proteomic analyses, we detected at least four different BARP isoforms in infected saliva (Tb927.9.15520, Tb927.9.15530 and Tb927.9.15570), in agreement with a previous report [34]. MISP were previously included in the trypanosome surface phylome [40], within the Fam50 family that interestingly also includes BARP, and the T. congolense proteins GARP [61,62] and CESP [63]. Apart from T. vivax, genomes from most of the disease-relevant species of African trypanosomes such as T. brucei (including T. b. gambiense and T. b. rhodesiense), T. congolense, and T. evansi encode misp homolog genes. The preservation of MISP only across trypanosome species with full development in the tsetse vector suggests a conserved function and potential role in development or transmission.

We have confirmed that misp and barp gene expression is developmentally regulated in T. b. brucei SG stages as previously determined [36,37,64]. Our semi-quantitative RT-PCR method has allowed the detailed study of individual misp paralogs by exploiting the different number of repetitive elements at their C-termini. We found that misp up regulation is not homogeneous across isoforms, but instead biased towards a preferential expression for the shortest misp-B paralogs. Except for misp-440 that seems to have a low but stable expression, misp-A transcripts are only detected during the late stages of parasite development in SG. It is unknown how these genes are expressed in such different ways despite being part of a gene array that is polycistronically-transcribed with >99.7% sequence conservation in UTR and intergenic sequences. Differences in homolog transcript levels, however, may not reflect changes in the respective protein abundance, which remain unknown as the polyclonal antibodies used in this study are expected to cross-react with all isoforms due to the overall high amino acid sequence conservation. Although proventricular trypanosomes produce large amounts of misp and barp transcripts [21], these are only translated into proteins in SG stages and localize evenly across the cell surface. While BARP expression peaks in SG epimastigotes and is gradually reduced during metacyclogenesis, MISP expression seems to increase with development and peaks in metacyclic forms, leading us to consider MISP as predominant metacyclic surface proteins (S21 Fig). Our observations disagree with a previous study [38] describing MISP (therein named SGE1) as proteins exclusively expressed by SG epimastigotes and absent in metacyclic forms, based on single-cell RNA-seq and immunostaining analyses. Such divergences may be attributed to several factors, including (1) use of different T. b. brucei strains and tsetse infection timing; (2) RNA-seq data not correlating with actual transcript translation like in PV-associated trypanosomes presenting high levels of MISP and BARP transcripts but not proteins; and (3) use of different polyclonal antibodies, particularly relevant when having to bind to scarcely accessible epitopes within the mVSG coat in metacyclic cells. Instead, we used confocal fluorescence microscopy and HRSEM to confirm the surface localization of MISP in metacyclic cells, and importantly, demonstrated that MISP epitopes are susceptible to antibody binding using live cells. Lastly, two more recent studies also support the expression of MISP in metacyclic cells using scRNA-seq [65], or a combination of transcriptomics and proteomics [66] (S3 File). Of importance, these studies were performed using either EATRO1125 [65] or Lister 427 (29:13 clone) parasite strains [66], which together with our MISP immunodetection in TSW196 (S15 Fig), strongly suggest that MISP expression in AnTat 1.1 metacyclic trypanosomes is not strain-specific. As for BARPs, previous observations pointed at their expression being lost in MCFs [20,23], but we detected significant amounts of transcripts and protein levels in MCFs with variable localization (2C Fig). A similar effect was observed in PCFs where ectopic MISP localized to the flagellum, but then regained surface localization when cells differentiated into MCFs (where MISP are natively expressed).

Metacyclic forms are cells with an arrested cell cycle with low transcription activity. Therefore, detecting high levels of misp transcripts prompted us to hypothesize that these proteins may be abundant and play an important role in parasite development, transmission, or in establishing a mammalian host infection. To test this, we attempted to determine misp essentiality by knocking down transcripts from all paralogs using a single RNAi construct due to the high sequence conservation. The developmental expression of misp only in SG stages limited the selection process of clones in the culturable stages of the parasite (PCF). Although partial downregulation was successfully achieved in salivary gland trypanosomes, MISP could still be detected in MCFs making results inconclusive despite the absence of phenotype. We then knocked out all ten misp alleles using CRISPR-Cas9, which did not result in any apparent phenotype in vitro, and only produced a slight reduction in salivary gland infectivity, partially supporting the findings obtained by RNAi. Altogether, these data suggest that MISP are not essential for the development of T. brucei in the vector and may function during or after transmission to a mammalian host [26]. It remains to be determined whether MISP subfamilies have either distinct or complementary functions, as adding back a single misp did not rescue the slight infection phenotype observed, which may be simply due to loss of virulence from the extended culture time needed to generate the transgenic cell lines.

To gain more insight into the function of MISP, we solved the crystal structure of the ordered N-terminal domain of the MISP360 (Tb427.07.360). Structural analysis revealed a triple helical bundle structure that resembles other trypanosome surface proteins such as the Tb/TcHpHbR [67], GARP [48] or a VSG monomer [47]. However, it is important to consider that mature MISP360 contains an additional C-terminal domain predicted to be highly disordered. In the absence of definitive structural data, we modelled the C-terminus, which indicated that it might serve as a long tether effectively elevating the N-terminal head group from the surface of the parasite membrane. This may partly explain the recognition of MISP by polyclonal antibodies shown by live immunostaining. It is tempting to speculate that this extension facilitates a biologically relevant interaction between the parasite and its environment such as promoting the acquisition of nutrients, serving as an adhesion or decoy, or modulating a response in the vertebrate host. The previously characterized TbHpHbR lies partially within the VSG layer of the BSF and allows the trypanosome to acquire nutrients from the blood of the mammalian host. The mechanism by which it acquires nutrients is facilitated by a C-terminal extension that increases its relative distance from the membrane, thus making the ligand-binding site accessible by protruding above the VSG layer [45]. Since MCFs express mVSGs [68] alongside MISP, a C-terminal extension could support a similar role although the substantial length of the extension would likely render cellular uptake of a nutrient payload difficult. However, the unstructured extension is also suited to support an adhesion role for MISP, similar to the function described for the Haemophilus Cha adhesins [69]. Intriguingly, the only study that defined an adhesin function of a trypanosome surface protein (T. congolense CESP) is also part of the Fam50 family of proteins [63], thereby suggesting that MISP could also play a role in adhesion. Our structural analyses revealed the existence of a predominantly hydrophobic pocket in the apical end of MISP360, which we suggest may allow the coordination of a molecular partner, consistent with the putative role as an adhesin. Despite high sequence identity between isoforms, the residues forming and surrounding the pocket lacked sequence conservation and resulted in a distorted cavity between the homolog models that may reflect the capacity to coordinate a structurally diverse set of ligands, potentially from different hosts. Furthermore, the homologs differ in the length of the C-terminal region, originating from the variable number of 26-residues motifs. This region could offer the parasite a robust mechanism to engage a variety of cell types or provide a sequential binding effect to enable tighter adhesion. It is noteworthy that these putative unstructured C-terminal regions contain lysine and arginine residues, which are inherently susceptible to proteolysis consistent with predictions using the PROSPER software [70] (S22 Fig). The predicted proteolytic susceptibility may serve a strategic function in enabling the parasite to release the molecular interactions anchoring it to the tsetse salivary glands and thereby promote transmission. It has been assumed that the MCF surface is structurally and functionally equivalent to that of BSFs as they show the same thick VSG (electron-dense) layer in TEM sections [12,24,71]. This implies that mVSGs also form a dense macromolecular barrier that possibly masks MCF invariant surface proteins from the host’s immune system. Most of the characterized BSF invariant surface proteins (e.g., ISG65, ISG75, ESAG4 and hexose transporters), which are found on either the flagellar or cell surface, are attached via transmembrane domains [8,11]. Two exceptions are the transferrin [9] and the HpHbR receptors [10], which are GPI-anchored and exclusively localize in the flagellar pocket. This restricted localization has been proposed to be due partly to the presence of only one GPI anchor molecule, differing from VSGs, which form homodimers (i.e. two GPI anchors) that allow their homogeneous surface expression [72]. In contrast, our structural and immunostaining evidence shows that MISP appear to be displayed on the entire cell surface as monomer. This may suggest that, unlike BSFs, MCFs allow the homogeneous surface expression of monomeric GPI-anchored invariant proteins, although we cannot rule out the possibility of MISP multimerizing in vivo. This clearly differs when MISP is ectopically expressed in BSF or PCF, in which the proteins localize at the flagellar pocket and/or the flagellar membrane. It remains to be determined the mode of association to the metacyclic membrane of another small family invariable surface proteins known as Fam10 [40], as these molecules lack identifiable hydrophobic transmembrane domain or GPI-signal addition site in the C-terminus.

Due to its surface exposure, antigenicity, and conservation, we considered MISP as potential vaccine candidates. Our immunization experiment failed to provide protection against a tsetse challenge, as it also failed against T. b. brucei needle challenge in a previous study [38]. However, in neither study was the MISP immunogen designed for optimal performance; improvements could include: 1) using an alternative expression system (e.g. Leishmania tarantolae) to conserve all post-translational modifications; 2) immunizing with full length protein; or 3) testing different adjuvants that elicit a protective immune response against a trypanosome infection.

Our proteomic analyses identified peptides from several VSG proteins. The identification of mVSG peptides in infected saliva is not surprising given that infected SGs contain thousands of quiescent metacyclic cells, which may release mVSGs into the SG lumen. mVSG release may occur partly by the action of the parasite GPI-PLC, which is highly expressed in MCFs [23,73] and/or by the action of tsetse salivary proteases. Interestingly, among the VSG species detected, only one corresponds to a canonical mVSG (mVAT4), which has not been previously detected as protein, but its metacyclic expression site (MES) has been well characterized [74]. Notably, a similar study detected mVAT5 in tsetse saliva infected, but from a different T. brucei strain [34]. Although it has been reported that BSFs may express mVSGs from bloodstream expression site (BES) at low levels [75], no study has described the expression of BSF VSG proteins in MCFs. This supports the assumption that, although recombination may occur from MES to BES, it does not appear to occur in the opposite direction since MES lack the 70 bp repeats upstream of the mVSG gene. However, we identified peptides found in at least four BSF VSGs, including MiTat 1.2 (identified by three unique peptides). This suggests that either (1) MCF could have active BES (unlikely since MCFs do not appear to express ESAG proteins [73]) or 2) there is recombination at the MES replacing the mvsg by a vsg gene, or 3) there is recombination between mvsg and vsg genes leading to the formation of mosaics. In addition, we cannot rule out that the undetermined VSG repertoire of the TSW196 strain [76] (used in this study) could differ from other reference strains [77]. Importantly, previous studies have found similar results although this aspect has been overlooked [34,36,64,73] (S6 Table). For example, at least two BSF VSGs were identified in saliva from flies infected with T. b. brucei EATRO 1125 [34]; one BSF vsg was found to be highly transcribed in a T. b. brucei RUMP503 SG infection [64]; and 15 BSF vsgs had similar transcription levels than mvsgs (and a few were detected as protein) in T. b. brucei Lister 427 MCF overexpressing the RNA-binding protein 6 [73].

In summary, not only does the T. brucei-infected tsetse saliva contain infective MCFs, but also a cocktail of G. morsitans salivary proteins, bacterial and trypanosome soluble factors. How this complex composition modulates parasite transmission is a fascinating question yet to be addressed. Among the detected parasite proteins in infected saliva, MISP represent a family of immunogenic, invariant surface proteins that are exposed on the cell surface of the mammalian-infectious metacyclic stage of T. brucei. Based on their external exposure and sequence conservation, we considered MISP as potential vaccine candidates against T. brucei spp-caused trypanosomiasis, in a similar way to the malaria vaccine that targets the sporozoite CSP protein [78]. Lastly, it could be advantageous to use MISP as a xenodiagnosis marker in the field for the identification of disease-transmissible tsetse. Having a parasite marker that discerns between a mature tsetse salivary gland infection (disease vector) and trypanosomes in the midgut (potential non-vectors), would provide a key tool for accurately assessing both vector incrimination and the prevalence of trypanosomiasis in the field.

Methods

Ethics statement

All applicable international, national, and institutional guidelines for the care and use of animals were followed. Specifically, the experimental use of the BALB/c female mice for the immunization and tsetse fly-mediated trypanosome infection was approved by the Institute of Tropical Medicine Antwerp (ITM) Animal Ethics Committee Clearance under licence No. ECD VPU2018-1.

T. b. brucei cell culture

Trypanosoma brucei brucei strain AnTat 1.1 90:13 PCF [55,79] were cultured in SDM-79 medium supplemented with 10% FBS, haemin, 2 mM L-glutamine, Glutamax (Gibco), 15 μg/mL G418 (InvivoGen) and 50 μg/mL hygromycin B (InvivoGen), at 27°C and 5% CO2. Cells were typically passaged every 2 days. Transgenic AnTat 1.1 90:13 PCF cell lines were cultured in the presence of the selection marker drugs puromycin 1 μg/mL or zeocin 10 μg/mL (InvivoGen). Cells requiring the activation of the transgene were cultured in the presence of fresh 1μg/mL doxycycline (in 70% ethanol) or the same volume of 70% ethanol for negative controls. T. b. brucei AnTat 1.1 90:13 BSF were cultured in HMI-9 medium supplemented with 10% FBS, 10% Serum Plus (Sigma) and GlutaMax (Gibco), 2.5 μg/mL G418 and 4 μg/mL hygromycin B at 37°C and 5% CO2.

Tsetse fly infection and dissection

G. m. morsitans were obtained from the tsetse insectary at the Liverpool School of Tropical Medicine, where they were maintained at 26°C (+/-1°C) and 68–78% relative humidity. Flies were fed for 12 minutes every second day on sterile, defibrinated horse blood (TCS Biosciences) through a silicone membrane feeding system. Teneral (24h post emergence, unfed) males were infected by combining T. b. brucei TSW196-infected rat blood with fresh, sterile, defibrinated horse blood at a density of 106 cells/mL (for saliva proteomics, S1 Fig). Alternatively, flies were infected by mixing either T. b. brucei AnTat 1.1 90:13 [55] or AnTat 1.1 Cas9 [80] PCF cells with washed, defibrinated horse blood at a density of 5x105 cells/mL. Uninfected control flies were fed on the same blood (free of trypanosomes) for the same amount of time. Any experimental flies that had not fed were removed the following day. Experimental flies were maintained for 4 weeks, feeding on normal defibrinated horse blood every 2 days. Alternatively, flies infected with mutant trypanosomes requiring tetracycline activation of the transgene were fed with normal defibrinated horse blood supplemented with fresh 2 μg/mL doxycycline (in 70% ethanol). Bloodmeals for control flies were supplemented with the same volume of 70% ethanol only. Flies were dissected at 28 to 30 days post infection, 72 hours after receiving the last bloodmeal. For saliva proteomics, based on normal T. b. brucei TSW196 salivary gland infection rates (~15%) and yields (~104 trypanosomes/gland pair), 1,500 naïve and 1,500 T. b. brucei-infected tsetse were dissected (3 biological replicates of ~500 male flies per group). Salivary glands were collected in sterile ice-cold PBS and centrifuged at 12,000 rpm for 5 minutes at 4°C to collect the resultant saliva supernatant. SG and saliva were then separately stored at -80°C until further analyses. Alternatively, the tsetse SG, PV and MG were dissected out (in this order to prevent trypanosome cross-contamination) in sterile ice-cold PBS and T. brucei cells were isolated, washed in PBS by centrifugation at 2,000 rpm for 5 minutes at 4°C and either stored at -80°C or directly used for immunostaining.

High-resolution nLC-MS/MS proteomic analysis

Saliva samples from naïve and T. brucei-infected tsetse were lyophilized and sent to the Dundee University Fingerprints Proteomics Facility for in-solution trypsin digests and identification by mass spectrometry. Briefly, lyophilized saliva samples (10–100 μg each) were redissolved in 200 μL 0.4 M NH4HCO3 containing 8 M urea, and disulfide bonds were reduced with 5 mM dithiothreitol for 15 min at 50°C. Cysteine residues were alkylated with freshly prepared 10 mM iodoacetamide for 30 min at room temperature (RT), protected from light. Proteins were digested with 20 μg sequencing-grade trypsin (Sigma-Aldrich) for 24 h at 37°C. The reaction was terminated by adding 10 μL formic acid. The resulting peptides were desalted in a C18-SPE cartridge (1 mL, Discovery DSC-18, Supelco, Sigma-Aldrich), and analyzed by nano-liquid chromatography-high-resolution tandem mass spectrometry (nLC-HR-MS/MS) using an Ultimate 3000 RSLC nano coupled to LTQ-Orbitrap Velos Pro MS (Thermo Fisher Scientific) and filtered with an estimated false-discovery rate (FDR) of 1%. The PD settings were as follows: HCD MS/MS; cysteine carboxyamidomethylation as fixed modification; methionine oxidation and lysine acetylation as variable modifications; fully tryptic peptides only; up to two missed cleavages; parent ion mass tolerance of 10 ppm (monoisotopic); and fragment mass tolerance of 0.6 Da (in SEQUEST) and 0.02 Da (in PD 2.1.1.21) (monoisotopic). Tandem MS/MS spectra were searched against the following databases: a combined protein database of T. b. brucei, T. b. gambiense and T. b. rhodesiense (downloaded in October 2017 from TriTrypDB; http://www.tritrypdb.org/ in FASTA format, release 34, 18,729 entries, and from UniProtKB; http://www.uniprot.org/; release 2017_10, 25,450 entries), and G. m. morsitans (downloaded in May 2017 from VectorBase; https://www.vectorbase.org/; version 1.7, 12,494 entries). Common contaminant sequences (MaxQuant) were included in this combined database. The resulting PD dataset was further processed through Scaffold Q+ 4.8.2 and perSPECtives (Proteome Software, Portland, OR). A protein threshold of 95%, peptide threshold of 95%, and a minimum number of one unique peptide was used for identification of proteins. Manual inspection of the MS/MS spectrum was always performed for identification of any protein based on a single peptide. Using BLAST2GO, proteins were BLAST searched against the non-redundant NCBI protein database and mapped and categorized according to gene ontology terms (GO). Sodalis glossinidius proteins were identified using the Liverpool School of Tropical Medicine in-house MASCOT MS/MS ion search using a protein database from UniProtKB generated from the latest re-annotated coding sequences in S. glossinidius [81], using the following settings: two missed trypsin cleavages, same peptide modifications described above, peptide tolerance of 1.2 Da, MS/MS tolerance of 1.2 Da and peptide charges set to +1, +2 and +3. The mass spectrometry proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE (www.ebi.ac.uk/pride/) partner repository with the dataset identifier PXD039684.

Bioinformatic analyses

All trypanosome sequences were obtained from TriTrypDB (www.tritrypdb.org) [82] or UniProt (www.uniprot.org) [83]. G. m. morsitans sequences were obtained from VectorBase (www.vectorbase.org) [84]. Sequence analyses and multiple alignments were done using Clustal Omega [85] and Geneious R9 (Biomatters Ltd). Alignment figures were made in Jalview [86]. Phylogenetic trees were generated using PhyML 3.1 [41] under default conditions with a set bootstrap of 500. Signal peptide, GPI-anchor signal peptide, and transmembrane domain predictions were determined using the servers SignalP 4.1 [87], PredGPI [88] and TMHMM 2.0 [89,90], respectively, under default settings. N-glycosylation predictions were run using the NetNGlyc 1.0 server (http://www.cbs.dtu.dk/services/NetNGlyc). MISP 3D protein modelling was performed using I-TASSER [5153] with the crystal structure of MISP360 as a structural template and multiple sequence alignments (Clustal Omega) with the query sequences, or with IntFOLD [54]. Structural homology searches were performed using DALI [91] and the conservation models using ConSurf [49,50].

PCR and semi-quantitative RT-PCR

Total RNA from cultured and tsetse-isolated trypanosome samples was extracted using the GeneJet RNA extraction kit (Thermo Scientific) following the manufacturer’s protocol. Extracted RNA was further treated with the TURBO DNAse kit (Ambion) to remove residual genomic DNA following the manufacturer’s instructions. Variable amounts of RNA were reverse-transcribed into cDNA at 52°C for 10 minutes using SuperScript IV (Invitrogen) and oligo-dT20 primer, inactivated at 80°C for 10 minutes and treated with E. coli RNAse H at 37°C for 20 minutes. The resultant RNA-free cDNA was used in a downstream PCR using the Taq 2x Mastermix (New England Biolabs) and the primer pairs (400 nM) 5’-GCCGAAAGGACGGCAGAGACG-3’ and 5’-GTGTATCCTCCTATAGATT CTGCATAGC-3’ for the detection of individual misp homologs (amplicons of 410 bp, 332 bp, 254 bp and 176 bp for misp360, misp380, misp440 and misp400/420, respectively); 5’-GGCTACTTTTGACAGTGTTATG-3’ and 5’-CTTCCTTCGCTTTCGTAACG-3’ to detect all barp homologs (270 bp amplicon). To determine misp knockdown levels in mispRNAi cells, the primers 5’-GATGCCTCTAAAGCGAGAGAAGG-3’ and 5’-GTACTGGCGGCACTATTGTCCTC-3’ were used (263 bp amplicon). Expression values were normalized against the expression of tert [92,93] using the primers 5’-AACAGTAAGTCCCGTGTGG-3’ and 5’-CAAAGTCGTACTTCAACATCAG-3’ (233 bp amplicon). For all RT-PCR reactions, we ran a non-template control, a control for DNAse treatment, for reverse-transcription and a positive control with 1 ng genomic DNA for standardization. PCR conditions were as follows: 95° for 2 minutes, followed by 30 or 35 cycles of 95°C for 30 seconds, annealing for 30 seconds at variable temperatures, and elongation at 68°C for 30 seconds. A last elongation step at 68° was performed for 2 minutes. The cycle number was set at the exponential phase of the reaction. All RT-PCR products were mixed with sample buffer and run in a 2% agarose gel at 100V for 50 minutes. Gels were stained with SYBR Safe (Invitrogen), imaged using a Bio-Rad Gel Doc EZ system and analyzed using the ImageLab software. Imaging conditions were kept identical for all measurements for comparison.

Generation of anti-MISP and anti-BARP antibodies

Anti-MISP polyclonal antibody was generated by immunization of a single rabbit (Davids Biotechnologie GmbH) with 3 courses of 0.3 mg of recombinant MISP380 (full Hisx6-tagged protein lacking N- and GPI-signal peptides produced in Rosetta 2 (DE3) E. coli and purified by nickel affinity followed by size exclusion chromatography), and further purification of IgG from antiserum (63 days post immunization bleed) by protein A and affinity (MISP) chromatography, and posterior exclusion chromatography using DE3 E. coli lysate). Anti-BARP polyclonal antibody was generated by immunization of a single rabbit (Eurogentec) with BSA conjugated to the epitope peptide sequence CKVQAEEAVELAESK (fully conserved across all BARP isoforms) and further purification of specific antibodies by protein A and affinity chromatography. We tested the specificity of either antibody using western blotting by probing recombinant MISP, BARP and soluble VSG221 (S23 Fig).

Immunofluorescence assay and confocal laser scanning microscopy

Cultured or tsetse-derived T. b. brucei cells were washed in PBS and fixed in fresh 4% paraformaldehyde (PFA) for 30 minutes at RT. Cells were then allowed to settle onto poly-L-lysine coated slides for 20 minutes in a humidity chamber. Alternatively, cells were first air-dried on a poly-L-lysine slide, fixed in 100% methanol for 10 minutes at -20°C and rehydrated in PBS after air-drying. PFA-fixed cells were then either permeabilized or not in 0.1% Triton X-100 for 10 minutes at RT. They were then blocked in vPBS (PBS containing 45.9 mM sucrose and 10 mM glucose) 20% FBS for 1 hour, washed in PBS and incubated for 1 hour at RT with the primary antibody solution; 1:200 (anti-MISP (this study), anti-BARP (this study), anti-HA (ThermoFisher), 1:50 (anti-CRD (Glyko)), 1:100 (anti-PFR (a generous gift from Prof Helen Price, Keele University), 1:800 (anti-EP (Cedar Lane)) dilutions in vPBS 20% FBS. After further washing, they were incubated for 1 hour at RT with the secondary antibody solution (1:500 goat-anti-mouse/rabbit Alexa Fluor 488/555; ThermoFisher). Cells were then incubated with 300 ng/mL 4’,6-diamidino-2-phenylindole (DAPI) for 10 minutes, washed and mounted in Slowfade Diamond mounting oil (Life Technologies). Slides were imaged and analyzed using the confocal laser scanning microscope Zeiss LSM-880 and the Zeiss Zen software. For mean intensity fluorescence measurements, images were taken conserving laser power, pinhole aperture and digital gain settings. Images were further analyzed using Zen and ImageJ.

High-Resolution Scanning Electron Microscopy (HRSEM)

Both T. brucei PCF and MCF cells were fixed in 4% PFA 0.02% glutaraldehyde in 0.1M phosphate buffer (PA) for 30 minutes at RT, washed in PBS and attached to poly-L-lysine coverslips. Cells were blocked in 1% BSA PBS for 30 minutes and incubated with primary anti-MISP antibodies (1:50 in blocking solution) for 1 hour at RT. After washing, cells were incubated with secondary IgG goat anti-rabbit conjugated to 10nm gold nanoparticles (1:50 in blocking solution) for 1 hour at RT. Cells were washed and fixed in 2.5% glutaraldehyde 0.1M PA for 30 minutes at RT, washed in water and dehydrated in 30, 50, 70, 90 and 100% ethanol solutions progressively, for 10 minutes each. Cells were finally dried using a Tousimis automatic critical point dryer and coated in 4nm palladium using an Agar Sputter Coater. Samples were imaged in a merlin VP compact high-resolution scanning electron microscope (Zeiss) equipped with a 3View stage (Ametek) and an OnPoint back-scattered electron detector (Ametek). The samples were imaged under vacuum with Nitrogen gas. The microscope conditions were as follows: aperture 20μm, working distance 5-6mm, Z height 47mm, EHT (accelerating voltage) 10kv, BSD of very high gain, brightness 46.9% and contrast 87.7%.

Generation of HA-eGFPMISP360 and mispRNAi PCF transgenic cell lines

To create a cell line expressing ectopic MISP360 tagged with an HA-epitope and eGFP at N-terminus (HA-eGFPMISP360), the nucleotide sequence coding for the mature Tb427.07.360 protein lacking the signal peptide (Asp18 to Phe382) was obtained by PCR (Q5 high-fidelity, New England Biolabs) from T. b. brucei (strain AnTat 1.1 90:13) genomic DNA using the following primer sequences: 5’-GCGTCTAGAGACTCCATAATTGAGGAAGG-3’ and 5’-GCGAGATCTTTAAAAATGTGCGGCAGC-3’ with XbaI and BglII restriction enzyme targets (underlined), respectively. The Tb927.7.360 signal peptide (Met1 to Ala17) coding sequence was separately obtained by annealing the oligonucleotides 5’-CGCAAGCTTATGACTAGCCCGTTTTTTGTGCTTGCTTGCTATACTTACGTATGTCACCGCCAAGCTTCGC-3’ and 5’- GCGAAGCTTGGCGGTGACATACGTAAGTATAGCAAG

CAAGCACAAAAAACGGGCAGTCATAAGCTTGCG-3’ with HindIII at both ends, and the HA epitope coding sequence was obtained by annealing the oligonucleotides 5’-CGCAAGCTTTATCCATATGACGTGCCGGATTATGCGACTAGTCGC-3’ and 5’-GCGACTAGTCGCATAATCCGGCACGTCATATGGATAAAGCTTGCG-3’ with HindIII and SpeI. The pDEX-577 plasmid [94] for the tetracycline-inducible ectopic gene expression under a T7 promoter, was digested with HindIII and SpeI and ligated with HindIII/SpeI-digested HA sequence. The resultant plasmid was digested with HindIII, dephosphorylated with recombinant shrimp alkaline phosphatase and the HindIII-digested signal peptide coding sequence was inserted. The construct was then digested with XbaI and BamHI to ligate the Tb927.7.360 (lacking the signal peptide) coding sequence digested with XbaI and BglII. The resultant construct was linearized using NotI high-fidelity (New England Biolabs) and transfected into T. b. brucei (strain AnTat 1.1 90:13) PCF using an Amaxa 4D nucleofector, program FI-115 in Tb-BSF buffer [95]. Clonal transfectant lines were selected with 10 μg/mL zeocin. To create a cell line for the RNAi knockdown of misp (mispRNAi), a short unique 494 bp sequence highly conserved in all misp homologs was manually selected and obtained by PCR (Q5 high-fidelity NEB) using the primers 5’-CTACCACTACAGCAATTCGG-3’ and 5’- GCCTCTCGCAATTCGTCTTG-3’ with the restriction target pairs XhoI/BamHI or NdeI/HindIII respectively. By digesting with the corresponding restriction enzymes, the two amplicons were sequentially inserted in sense and antisense directions into the plasmid pALC14 [55] for a tetracycline-inducible transcription of a short hairpin double-stranded RNA. The plasmid was linearized (NotI high-fidelity) and transfected into AnTat 1.1 90:13 PCF as described above, and clonal transfectant cell lines were selected with 1 μg/mL puromycin. Transgene insertions were checked by PCR.

Generation of MISP knockout and addback cell lines using CRISPR/Cas9

All ten misp alleles (5 paralogue genes) were knocked out in two consecutive rounds of CRISPR/Cas9 in AnTat 1.1 Cas9 as previously described [80], using the LeishGEdit toolkit [96], except for transfections being carried out in PCF cells obtained by in vitro differentiation from cultured BSF in DTM medium containing 6mM cis-aconitate at 27°C. Briefly, dsDNA encoding for two different sgRNA targeting universal sequences upstream and downstream the coding sequence of all misp paralogs, were produced by PCR and co-transfected as described above with two donor DNA, consisting of either neomycin or hygromycin resistance genes flanked by homology regions found beyond the CRISPR-targeted sites. The addback cell line was generated by complementing the MISP KO clone 3 cell line with a pDEX377 plasmid containing the misp360 gene for its tetracycline-inducible ectopic expression.

Expression and purification of recombinant MISP360 and BARP for structural analyses

The nucleotide sequence encoding Tb427.07.360 (MISP360) (Gly24 to Ala234) and BARP ABB49055.1 (Glu23 to Gly238) were synthesized by GenScript and codon-optimized for E. coli. The constructs were sub-cloned into a modified pET32a vector (Novagen) containing N-terminal thioredoxin (Trx) and hexa-histidine (His6) tags separated from the gene of interest by a TEV protease site. Sequencing confirmed that no mutations were introduced. E. coli Rosetta-gami 2 (DE3) cells (Novagen) were transformed with the plasmid and grown in autoinduction medium (Invitrogen) from a 5% inoculum. Following four hours of growth at 37°C and 64 hours at 16°C, cells were harvested, and the pellet re-suspended in 20 mM HEPES pH 8.3, 1 M NaCl, 30 mM imidazole and 5mM β-mercaptoethanol. Cells in suspension were lysed using a French Press, insoluble material was removed by centrifugation, and soluble fraction allowed to batch-bind with Ni-agarose beads at 4°C for 1 hour. Recombinant MISP360 and BARP were eluted with the same buffer as before supplemented with 250 mM imidazole, and fractions were analyzed by SDS-PAGE and pooled based on purity. The Trx-His6 tag was removed by TEV cleavage overnight at 291 K, and the proteins were further purified by size exclusion chromatography (SEC) (Superdex 16/60 200) in HEPES buffered saline (HBS) with 1mM DTT. Protein concentration was determined by absorbance at 280 nm with calculated extinction coefficients of 10220 M-1 cm-1 (MISP360), and 12740 M-1 cm-1 (BARP).

Crystallization and data collection of MISP360

Crystals of purified MISP360 were initially identified in the Index screen (Molecular Dimensions) using the sitting drop method at 18°C. The final drops consisted of 0.3 μL of MISP360 at 20 mg ml-1 with 0.3 μL of reservoir solution and were equilibrated against 60 μL of reservoir solution. Diffraction-quality crystals grew in 48 hours in 25% PEG, 1500. A single crystal was looped, cryopreserved in 12.5% glycerol for 20 seconds, and flash-cooled to 100 K directly in the cryostream. Diffraction data were collected on beamline 9–2 at Stanford Synchrotron Radiation Laboratory (SSRL) at a wavelength of 0.9795 Å.

MISP360 structure data processing, solution, and refinement

Diffraction data to 1.82 Å resolution were processed using Imosflm [97] and Scala [98] in the CCP4 suite of programs [99]. Initial phases were obtained by molecular replacement using PHASER [100] with one copy of GARP [48] (PDB 2Y44; 15% identity over 210 residues). Solvent molecules were selected using COOT [101] and refinement carried out using Phenix Refine [102]. The overall structure of MISP360 was refined to an Rfree of 20.43%. Stereo-chemical analysis performed with PROCHECK and SFCHECK in CCP4 [99] showed excellent stereochemistry with more than 99% of the residues in the favored conformations and no residues modelled in disallowed orientations of the Ramachandran plot. Overall, 5% of the reflections were set aside for calculation of Rfree. Data collection and refinement statistics are presented in S7 Table. The atomic coordinates and structure factors have been deposited in the Protein Data Bank under the code 5VTL.

In silico homology modelling of BARP

For modelling BARP, the crystal structure of GARP (PDB: 2Y44) and MISP360 were used as the templates, with 24 and 28% identity, respectively. Using Modeller 9v18 [43], 10 models of BARP were generated and the best model was chosen based on the low value of normalised discrete optimized protein energy (DOPE). The assessment of the final model was carried out with Ramachandran statistics [103], QMEAN [104], and ProSA [105].

SDS-PAGE and western blotting

Trypanosome lysates were prepared by pelleting 106 cells and heat-denaturing (98°C for 10 minutes) in Laemmli buffer. Samples were resolved on a 12.5% SDS-PAGE gel using the Bio-Rad mini-Protean system at 50V for 30 minutes followed by 2 hours at 120V. Proteins were transferred from the gel to a polyvinylidene fluoride (PVDF) membrane. After transfer at 90V for 1 hour, membranes were blocked (PBS-Tween 0.5% v/v, 5% w/v skim milk powder) for 1 hour at RT and incubated with primary anti-HA (1:10,000; Thermofisher) or anti-Sodalis Hsp60 mAb 1H1 [106] (mouse anti-symbiont GroEL; 1:10 dilution) overnight at 4°C. The membranes were then washed in PBS-Tween (0.5% v/v) and probed at RT for 1 hour with secondary goat-anti-rabbit or goat-anti-mouse IgG horseradish peroxidase-conjugated antibodies (1:25,000 dilution; Thermofisher). The membranes were washed and probed with Super Signal West Dura substrate (Thermo Scientific) and the signal transferred onto X-ray Amersham Hyperfilm (GE Healthcare). Membranes were lastly stained in 0.2% (w/v) nigrosine in PBS for 15 minutes for loading control. Alternatively, recombinant protein samples (S23 Fig) were denatured in LDS sample buffer with DTT reducing agent for 10 minutes at 70°C, run in precast 4–12% Bolt SDS-PAGE gels (Thermofisher) in MES buffer, dry-transferred onto PVDF membranes, blocked with 1% BSA in TBS, and incubated with primary antibodies for 1 hour at RT (anti-MISP 1:2,000; anti-BARP 1:2,000; anti-VSG 1:5,000). After washing, membranes were incubated with secondary Alexa Fluor Plus 800 anti-rabbit IgG antibodies (1:10,000, Thermofisher) for 1 hour at RT before imaging the membrane using an Oddysey Fc Imager.

MISP and BARP vaccine formulation, mouse immunization and tsetse challenge

For immunization, the N-terminal ectodomain of MISP360 (Asp18-Glu237) was recombinantly expressed in E. coli Rosetta 2 (BL21) cells (Merk), fused to a 6xHis-tag at C-terminus. The recombinant BARP used in immunization was identical to that used for crystallization (see above). Recombinant proteins were purified through a Ni-NTA column and SEC-FPLC and depleted from endotoxins using the Pierce high-capacity endotoxin removal resin (Thermo Scientific). Endotoxin levels were then determined using the Pierce Chromogenic endotoxin quant kit (Thermo Scientific). Vaccines were formulated in liposomal-monophosphoryl-Lipid A (L-MPLA) as previously described [107,108]. Briefly, 1,2-dimyristoyl-sn-glycero-3-phosphocholine, cholesterol, and 1,2-dimyristoyl-sn-glycero-3-phospho-rac-(1-glycerol) (Avanti Polar Lipids, Alabaster, AL) were dissolved in freshly distilled chloroform stabilized with 0.75% ethanol (9:7.5:1 molar ratio) along with MPLA (Sigma-Aldrich, St. Louis, MO) for a final concentration of 100 μg/mL. Liposomes were dried under nitrogen stream and placed overnight in a desiccator containing silica-gel desiccant beads to obtain a white pellet (Thermo Fisher Scientific), and then vortexed to resuspend at 10 μg/100 μL Lipid A per dose using sterile PBS (pH 7.4) containing either 5 μg of rMISP360, 5 μg of rBARP, 105 whole frozen metacyclics T. b. brucei Antat 1.1 in PBS, or PBS alone until completely dissolved. Vaccines were filtered (0.2-μm) before immunization and administered to groups of ten BALB/c 6–8 weeks old female mice three times at days 0, 10 and 21 by sub-cutaneous injection at the base of the tail. Pre-immune sera were collected for each individual 5 days before the first vaccine was delivered. Mice were then anaesthetized with ketamine (100mg/kg) and xylazine (20mg/kg) and challenged 12 days after the last booster (day 33) with the bite of tsetse flies pre-selected for positive T. b. brucei AnTat 1.1 salivary gland infections, by allowing individual tsetse to bite on the mouse ear dermis. Parasitaemia development was monitored by microscopic analysis using Uriglass disposable slides (Menarini Diagnostics) from day 4 post-challenge onwards. The tsetse flies used for this challenge were infected by providing a first blood meal after emergence (<48 hours) containing 106/mL T. b. brucei AnTat 1.1 pleiomorphic bloodstream forms in defibrinated horse blood supplemented with 10 mM reduced L-glutathione to increase trypanosome infection establishment. Flies were then maintained by feeding every 2–3 days on uninfected defibrinated horse blood. Salivary gland infected flies were pre-selected by induced probing on pre-warmed glass slides that were microscopically examined for the presence of metacyclic trypanosomes. To confirm the efficacy of vaccines in raising a specific IgG response against the antigens, we performed western blotting on recombinant MISP, BARP and metacyclic cell lysates probing with mouse antisera collected at day 30 post-vaccination (3 days pre-challenge). Briefly, samples were denatured in sample buffer and run in 10% SDS-PAGE gels, transferred onto nitrocellulose membranes, blocking in TBS-T 5% skim milk blocking buffer and probed with mouse anti-serum (1:200 dilution in blocking buffer) during 1 hour at RT. Membranes were then probed with anti-mouse IgG conjugated with HRP (1:20,000 dilution in blocking buffer) at 1 hour at RT, and antibodies detected using TMB-membrane peroxidase substrate (Seracare).

Supporting information

S1 Fig. Workflow for the analysis of T. b. brucei-infected tsetse saliva.

For each biological replicate, saliva samples from naïve flies were also collected and processed in parallel.

https://doi.org/10.1371/journal.ppat.1011269.s001

(TIF)

S2 Fig. Chemiluminescent Western blot analysis of G. m. morsitans naïve and T. b. brucei-infected saliva.

Mouse anti-Sodalis monoclonal antibody 1H1 recognizes a heat shock protein (Hsp60) produced by Sodalis glossinidius. Lanes 1 and 2 represent film exposed for 10 seconds. Lanes 3 and 4 show the nigrosine-stained PVDF membrane used for blotting to visualize the protein loading in each lane. Naïve saliva (N). Infected saliva (I).

https://doi.org/10.1371/journal.ppat.1011269.s002

(TIF)

S3 Fig. Multiple protein sequence alignments of MISP.

A, Multiple protein alignment of all MISPs (T. congolense MISP excluded). Residues colored based on identity conservation from black (maximum) to red (minimum). Alignment made using a Clustal Omega BLOSUM62 matrix with a gap open cost 10 and a gap extension cost of 0.1, using Geneious R9. Domains highlighted: ER signal peptide (green), potential N-glycosylation sites (yellow asparagines), C-terminus motifs (blue), and GPI-anchor attachment signal peptide (red). B, Multiple protein alignment of T. congolense MISP with MISP360 (MISP-A) and MISP400 (MISP-B). Domains highlighted as in ‘A’.

https://doi.org/10.1371/journal.ppat.1011269.s003

(TIF)

S4 Fig. Semi-quantitative RT-PCR expression analysis of misp mRNA in different trypanosome stages.

BSF, PCF and SG samples shown as in Fig 2A. The HA-GFPMISP360 PCF line either uninduced (360 -) or tet induced (360 +) was introduced expression control. The inset is a representative image of a DNA agarose gel displaying 4 bands corresponding to the RT-PCR amplification of the different misp homologs. Expression of the different misp isoforms was normalized to the expression of tert. Error bars represent +S.D. from two technical replicates (n = 2).

https://doi.org/10.1371/journal.ppat.1011269.s004

(TIF)

S5 Fig. Multiple protein alignment of the 14 BARP homologs in T. b. brucei.

Residues colored based on identity conservation from black (maximum) to red (minimum). Alignment made using a Clustal Omega BLOSUM62 matrix with a gap open cost 10 and a gap extension cost of 0.1, using Geneious R9. The detected BARP peptides in the nLC-MS/MS analysis are highlighted in blue. Domains highlighted: ER signal peptides (green), potential N-glycosylation sites (yellow asparagine residues) and GPI-anchor attachment signal peptides (red).

https://doi.org/10.1371/journal.ppat.1011269.s005

(TIF)

S6 Fig. Representative immunostaining images.

Non-infected salivary gland (Control); a) Tbb-infected (b-e) salivary glands immunostained with either anti-MISP (cyan) or anti-BARP (yellow) polyclonal antibody; DAPI DNA counterstain (magenta); scale bars 200 μm (a, b, c) and 20 μm (d, e).

https://doi.org/10.1371/journal.ppat.1011269.s006

(TIF)

S7 Fig. Representative negative controls images.

T. b. brucei AnTat 1.1 cultured bloodstream forms (BSF), procyclic cultured form (PCF), midgut procyclic form (PF), proventricular mesocyclic form (MSC), and metacyclic form (MCF) immunostained with anti-MISP polyclonal antibody (cyan); DAPI (magenta); merged and differential interference contrast (DIC); scale bars = 5 mm.

https://doi.org/10.1371/journal.ppat.1011269.s007

(TIF)

S8 Fig. Expression of EP-procyclins in proventricular trypanosomes.

Parasites extracted from infected PV at 30 d.p.i. Mesocyclics (MSC), long epimastigote forms (LEMF) and short epimastigote forms (SEMF) probed with anti-EP monoclonal antibody (yellow) and DAPI (blue). Nuclei (N) and kinetoplastids (K) noted in the blue channel. Scale bars: 5 μm.

https://doi.org/10.1371/journal.ppat.1011269.s008

(TIF)

S9 Fig. Mean relative fluorescence intensity of anti-body staining.

Anti-MISP (blue) and anti-BARP (orange) staining in attached epimastigote forms (EMF) and metacyclic forms (MCF); bars represent mean values; dots represent individual cell values. error bars indicate ±S.D.

https://doi.org/10.1371/journal.ppat.1011269.s009

(TIF)

S10 Fig. Trypanosome populations in infected salivary glands.

Cell type distribution (percentage) in SG-extracted parasites used for the quantification of mean fluorescence intensities of MISP (A) and BARP (B). Metacyclic forms (MCF), pre-metacyclic forms (P-MCF), epimastigotes with variable kDNA (K) and nuclei (N) numbers (E(1K1N), E(2K1N), E(2K2N)), multi-nucleated epimastigotes (E(multi)) and abnormal cells. C, Distribution (percentage) of types of localization of MISP and BARP (D), either flagellar (flagellum) or on the whole cell surface (surface).

https://doi.org/10.1371/journal.ppat.1011269.s010

(TIF)

S11 Fig. Characterization of salivary gland forms.

A, Schematic of a metacyclic and an epimastigote cell indicating the measurements in ‘B’: total cell length (red), nucleus to posterior end (orange), nucleus to kDNA (green), and kDNA to posterior end (black); anterior end (A), posterior end (P), kDNA (K), nucleus (N). B, Length measurements (mm) described in ‘A’ in metacyclic (n = 30; gray) and epimastigote cells (n = 30; blue) isolated from tsetse infected salivary glands (from 3 different biological replicates); circles indicate individual data points; thick horizontal line indicates mean value; error bars indicate ±SD; asterisks indicate significance (*** p<0.001); ns indicates non-significant differences; from one-sided t-test assuming normal distribution.

https://doi.org/10.1371/journal.ppat.1011269.s011

(TIF)

S12 Fig. Indirect detection of mVSGs in T. brucei SG stages.

Immunostaining of epimastigotes (EMF), pre-metacyclics (P-MCF) and metacyclics (MCF) from tsetse infected SG with polyclonal anti-CRD antibody. Nuclei (N) and kinetoplastids (K) noted in the magenta channel. Differential interference contrast (DIC), DAPI (magenta) and anti-CRD (blue). Scale bars: 5 μm.

https://doi.org/10.1371/journal.ppat.1011269.s012

(TIF)

S13 Fig. Ectopic localization of tagged MISP360.

A, DNA construct encoding for the ectopic tagged MISP360. B, Immunoblotting to detect ectopic (HA-tagged) MISP360 expressed by the mutant HA-GFPMISP360 PCF cell line, either with the transgene uninduced (-) or tet-induced (+), probed with anti-HA (top). PVDF membrane stained with nigrosine after film exposure for sample loading control. C, Cellular localization of the HA-tagged ectopic MISP360 in AnTat 1.1 90:13 procyclic (PCF) and metacyclic form (MCF), detected by immunostaining on PFA-fixed non-permeabilized PCF cells with either anti-HA plus anti-MISP (co-localization of ectopic tag with MISP), or with anti-HA plus anti-PFR (flagellar marker) in methanol-fixed cells. Scale bars: 5 μm.

https://doi.org/10.1371/journal.ppat.1011269.s013

(TIF)

S14 Fig. Immunodetection of MISP in live SG trypanosomes.

Representative images of live immunostaining of SG parasites using anti-MISP. Epimastigote (EMF) and metacyclic form (MCF) stained with anti-MISP (cyan) and DAPI (magenta). Nuclei (N) and kinetoplasts (K) noted in the blue channel. Scale bars: 5μm. DIC: differential interference contrast.

https://doi.org/10.1371/journal.ppat.1011269.s014

(TIF)

S15 Fig. Immunodetection of MISP expression in TSW196 stain trypanosome.

Representative metacyclic cell (MCF) obtained from infected tsetse salivary glands immunostained with polyclonal anti-MISP antibody (cyan); DAPI (magenta); merge, and differential interference contrast (DIC); scale bar = 5 μm. TSW-196 epimastigote and pre-metacyclic cells were also detected by the anti-MISP antibody. Salivary gland stages obtained from three independent tsetse infections with TSW-196 strain were performed and successfully immunostained with anti-MISP polyclonal antibody.

https://doi.org/10.1371/journal.ppat.1011269.s015

(TIF)

S16 Fig. Purification of recombinant MISP and BARP proteins.

A, SDS-PAGE (Coomassie blue-stained) of SEC elution fractions of the recombinant MISP360 ectodomain (ladder displaying relative molecular weight on the left (kDa). B, Complete SDS-PAGE with fraction of rMISP360 used to determine its crystal structure. C, Size exclusion chromatogram of rBARP used for crystallography (peaks for free TRX tag and rBARP monomer indicated). D, SDS-PAGE (Coomassie-stained) of rBARP fractions.

https://doi.org/10.1371/journal.ppat.1011269.s016

(TIF)

S17 Fig. Comparison of the crystal structure of T. brucei MISP360 N-terminus with other trypanosome surface proteins identified as structural analogues.

MISP360 (PDB: 5VTL), BARP (high confidence model), GARP (PDB: Y44), TcHpHbR (PDB: 4E40), TbHpHbR (PDB: 4X0J), VSG 221 monomer (PDB: 1VSG). Structures colored from blue (N-terminus) to red (C-terminus); molecular surface in semi-transparent grey.

https://doi.org/10.1371/journal.ppat.1011269.s017

(TIF)

S18 Fig. Suggested models of MISP C-terminal domains.

A, Height comparison between the crystal structure of VSG MiTat 1.2 N-terminus (PDB: 1VSG), the MISP360 N-terminus and the BARP model. B, Models of the five T. brucei MISP isoforms on the metacyclic surface. mVSGs modelled using the mVAT4 protein sequence and the structures of VSG MiTat 1.2 N- (PDB: 1VSG) and C-terminus (PDB: 1XU6). MISPs modelled using the crystal structure of MISP360 N-terminus.

https://doi.org/10.1371/journal.ppat.1011269.s018

(TIF)

S19 Fig. Characterization of misp RNAi cells.

A, In vitro growth curve of mispRNAi PCF cells, either induced (Dox+) or uninduced (Dox-). B, Relative misp RNA expression in mispRNAi PCF cells, midgut procyclics (MG), proventricular parasites (PV) and salivary gland forms (SG). Expression levels of Dox- cells are normalized to 100%. C, Representative immunostaining images of non-permeabilized mispRNAi MCF (Dox-/+) with anti-MISP (red), DAPI (blue) and phase. D, MISP mean fluorescence intensities (arbitrary units) on mispRNAi MCF (Dox-/+). E, Percentage of flies with mispRNAi cells infecting the midgut (MG), proventriculus (PV) and salivary glands (SG). Scale bars: 5μm. Error bars show standard deviation, asterisks represent significance (* for p-value < 0.05; *** for p-value < 0.001).

https://doi.org/10.1371/journal.ppat.1011269.s019

(TIF)

S20 Fig. Representative images T. b. brucei AnTat 1.1 wild type (WT), MISP knockout (KO), and MISP induced addback (AB) trypanosomes.

Cells were immunostained with either anti-CRD polyclonal antibody (A, green), or anti-BARP polyclonal antibody (B, yellow); DAPI (magenta), merged, and differential interference contrast (DIC); scale bars = 5 mm.

https://doi.org/10.1371/journal.ppat.1011269.s020

(TIF)

S21 Fig. Schematic of the expression of major GPI-anchored surface glycoproteins throughout the life cycle of T. brucei.

Bloodstream forms (BSF), procyclic forms (PF), mesocyclic forms (MSC), short (SE) and long (LE) epimastigotes infecting the proventriculus, attached epimastigotes infecting the salivary glands (EMF (SG)), pre-metacyclic forms (P-MCF), metacyclic forms (MCF). Representative immunostaining images of the parasite stages highlighting the nucleus and kinetoplast (magenta), and the cell surface (white) (top). Bottom bars define the duration of the expression of the surface proteins in relation to parasite developmental stage.

https://doi.org/10.1371/journal.ppat.1011269.s021

(TIF)

S22 Fig. PROSPER protease cleavage site predictions.

MISP360 (top) and MISP400 (bottom) amino acidic sequences with predicted protease cleavage sites highlighted (see key for protease types). Only one member of each MISP subfamily is represented.

https://doi.org/10.1371/journal.ppat.1011269.s022

(TIF)

S23 Fig. Characterization of anti-MISP antibody by western blotting.

A, Western blotting of soluble VSG 221 (sVSG), full recombinant MISP380 (rMISP), and recombinant BARP (rBARP) probed with either anti-MISP polyclonal antibody (left), or anti-BARP polyclonal antibody (centre), or anti-VSG 221 polyclonal antibody (right); ladder on the left for apparent molecular weights (kDa). B, Blotting membranes developed in ‘A’ stained with nigrosine. C, Instablue-stained SDS-PAGE analysis of same samples as in ‘A’.

https://doi.org/10.1371/journal.ppat.1011269.s023

(TIF)

S1 File. T. brucei BARP, VSG and MISP peptides identified by MS-MS/MS and corresponding supporting spectra.

https://doi.org/10.1371/journal.ppat.1011269.s024

(PDF)

S2 File. Vaccination experiment schedule and western blotting analyses of mouse antisera.

https://doi.org/10.1371/journal.ppat.1011269.s025

(PDF)

S3 File. Quantification of T. brucei MISP, BARP, mVSG, and alpha tubulin peptides identified by proteomics analyses in time course RBP6-overexpression experiments performed in Doleželová E. et al. (2020) [66].

https://doi.org/10.1371/journal.ppat.1011269.s026

(PDF)

S1 Table. List of Glossina morsitans morsitans proteins identified by proteomics analyses in either naïve or T. brucei-infected saliva.

https://doi.org/10.1371/journal.ppat.1011269.s027

(XLSX)

S2 Table. List of Sodalis glossinidius proteins identified by proteomics analyses in either naïve or T. brucei-infected saliva.

https://doi.org/10.1371/journal.ppat.1011269.s028

(XLSX)

S3 Table. List of Trypanosoma brucei brucei proteins identified by proteomics analyses in T. brucei-infected saliva.

https://doi.org/10.1371/journal.ppat.1011269.s029

(XLSX)

S4 Table. Intracellular T. brucei brucei proteins identified by proteomics analyses in T. brucei-infected tsetse saliva.

https://doi.org/10.1371/journal.ppat.1011269.s030

(DOCX)

S5 Table. In silico predictions of signal peptide, GPI-anchor peptide and N-glycosylation sites in all MISP isoforms based on amino acid sequences.

https://doi.org/10.1371/journal.ppat.1011269.s031

(DOCX)

S6 Table. Summary of T. brucei brucei VSG species found in either cultured or fly-derived metacyclics according to literature.

https://doi.org/10.1371/journal.ppat.1011269.s032

(DOCX)

S7 Table. Data collection and refinement statistics of TbMISP360 crystal structure.

https://doi.org/10.1371/journal.ppat.1011269.s033

(DOCX)

Acknowledgments

We are grateful to Douglas Lamont and Samantha Kosto (University of Dundee FingerPrints Proteomics Facility) for proteomic analysis and uploading data into PRIDE, and the Biomolecule Analysis and Omics Unit (BAOU) at UTEP/BBRC for performing part of the proteomic data analyses. The authors would like to thank the technical support staff at the Stanford Synchrotron Radiation Lightsource, Drs Federico Rojas and Keith Matthews for generous supply of T. brucei AnTat 1.1 Cas9 line, Dr Ian Hastings for statistical advice, and Dr Jay Bangs for kindly supplying aliquots of purified T. brucei VSG and anti-VSG antibody. We thank Drs Alena Zíková and Eva Doleželová for helpful discussions and kindly providing MISP expression data in RBP6 cells. We thank members of the Acosta-Serrano group for valuable input and, Dan Southern and Rob Leyland for maintaining the tsetse colony at LSTM.

References

  1. 1. Giordani F, Morrison LJ, Rowan TG, DEK HP, Barrett MP. The animal trypanosomiases and their chemotherapy: a review. Parasitol. 2016;143: 1862–89. pmid:27719692
  2. 2. Autheman D, Crosnier C, Clare S, Goulding DA, Brandt C, Harcourt K, et al. An invariant Trypanosoma vivax vaccine antigen induces protective immunity. Nature. 2021;595: 96–100. pmid:34040257
  3. 3. Matthews KR. The developmental cell biology of Trypanosoma brucei. J Cell Sci. 2005;118: 283–90. pmid:15654017
  4. 4. Horn D. Antigenic variation in African trypanosomes. Mol Biochem Parasitol. 2014;195: 123–9. pmid:24859277
  5. 5. Mugnier MR, Stebbins CE, Papavasiliou FN. Masters of Disguise: Antigenic Variation and the VSG Coat in Trypanosoma brucei. PLoS Pathog. 2016;12: e1005784. pmid:27583379
  6. 6. Pinger J, Nesic D, Ali L, Aresta-Branco F, Lilic M, Chowdhury S, et al. African trypanosomes evade immune clearance by O-glycosylation of the VSG surface coat. Nat Microbiol. 2018;3: 932–8. pmid:29988048
  7. 7. Engstler M, Pfohl T, Herminghaus S, Boshart M, Wiegertjes G, Heddergott N, et al. Hydrodynamic flow-mediated protein sorting on the cell surface of trypanosomes. Cell. 2007;131: 505–15. pmid:17981118
  8. 8. Paindavoine P, Rolin S, Van Assel S, Geuskens M, Jauniaux JC, Dinsart C, et al. A gene from the variant surface glycoprotein expression site encodes one of several transmembrane adenylate cyclases located on the flagellum of Trypanosoma brucei. Mol Cell Biol. 1992;12: 1218–25. pmid:1545803
  9. 9. Grab DJ, Shaw MK, Wells CW, Verjee Y, Russo DC, Webster P, et al. The transferrin receptor in African trypanosomes: identification, partial characterization and subcellular localization. Eur J Cell Biol. 1993;62: 114–26. pmid:8269969.
  10. 10. Vanhollebeke B, De Muylder G, Nielsen MJ, Pays A, Tebabi P, Dieu M, et al. A haptoglobin-hemoglobin receptor conveys innate immunity to Trypanosoma brucei in humans. Science. 2008;320: 677–81. pmid:18451305
  11. 11. Ziegelbauer K, Multhaup G, Overath P. Molecular characterization of two invariant surface glycoproteins specific for the bloodstream stage of Trypanosoma brucei. J Biol Chem. 1992;267: 10797–803. pmid:1587856.
  12. 12. Vickerman K. Developmental cycles and biology of pathogenic trypanosomes. Br Med Bull. 1985;41: 105–14. pmid:3928017
  13. 13. Grandgenett PM, Otsu K, Wilson HR, Wilson ME, Donelson JE. A function for a specific zinc metalloprotease of African trypanosomes. PLoS Pathog. 2007;3: 1432–45. pmid:17953481
  14. 14. Gruszynski AE, DeMaster A, Hooper NM, Bangs JD. Surface coat remodeling during differentiation of Trypanosoma brucei. J Biol Chem. 2003;278: 24665–72. pmid:12716904
  15. 15. Aksoy E, Vigneron A, Bing X, Zhao X, O’Neill M, Wu YN, et al. Mammalian African trypanosome VSG coat enhances tsetse’s vector competence. Proc Natl Acad Sci U S A. 2016;113: 6961–6. pmid:27185908
  16. 16. Roditi I, Schwarz H, Pearson TW, Beecroft RP, Liu MK, Richardson JP, et al. Procyclin gene expression and loss of the variant surface glycoprotein during differentiation of Trypanosoma brucei. J Cell Biol. 1989;108: 737–46. pmid:2645304
  17. 17. Acosta-Serrano A, Vassella E, Liniger M, Kunz Renggli C, Brun R, Roditi I, et al. The surface coat of procyclic Trypanosoma brucei: programmed expression and proteolytic cleavage of procyclin in the tsetse fly. Proc Natl Acad Sci USA. 2001;98: 1513–8. pmid:11171982
  18. 18. Vassella E, Den Abbeele JV, Butikofer P, Renggli CK, Furger A, Brun R, et al. A major surface glycoprotein of Trypanosoma brucei is expressed transiently during development and can be regulated post-transcriptionally by glycerol or hypoxia. Genes Dev. 2000;14: 615–26. pmid:10716949.
  19. 19. Rose C, Casas-Sanchez A, Dyer NA, Solorzano C, Beckett AJ, Middlehurst B, et al. Trypanosoma brucei colonizes the tsetse gut via an immature peritrophic matrix in the proventriculus. Nat Microbiol. 2020;5: 909–16. pmid:32313202
  20. 20. Sharma R, Gluenz E, Peacock L, Gibson W, Gull K, Carrington M. The heart of darkness: growth and form of Trypanosoma brucei in the tsetse fly. Trends Parasitol. 2009;25: 517–24. pmid:19747880
  21. 21. Urwyler S, Studer E, Renggli CK, Roditi I. A family of stage-specific alanine-rich proteins on the surface of epimastigote forms of Trypanosoma brucei. Mol Microbiol. 2007;63: 218–28. pmid:17229212
  22. 22. Van Den Abbeele J, Claes Y, van Bockstaele D, Le Ray D, Coosemans M. Trypanosoma brucei spp. development in the tsetse fly: characterization of the post-mesocyclic stages in the foregut and proboscis. Parasitol. 1999;118: 469–78. pmid:10363280
  23. 23. Rotureau B, Subota I, Buisson J, Bastin P. A new asymmetric division contributes to the continuous production of infective trypanosomes in the tsetse fly. Development. 2012;139: 1842–50. pmid:22491946
  24. 24. Tetley L, Turner CM, Barry JD, Crowe JS, Vickerman K. Onset of expression of the variant surface glycoproteins of Trypanosoma brucei in the tsetse fly studied using immunoelectron microscopy. J Cell Sci. 1987;87: 363–72. pmid:3654788
  25. 25. Ginger ML, Blundell PA, Lewis AM, Browitt A, Gunzl A, Barry JD. Ex vivo and in vitro identification of a consensus promoter for VSG genes expressed by metacyclic-stage trypanosomes in the tsetse fly. Eukaryot Cell. 2002;1: 1000–9.
  26. 26. Caljon G, Van Reet N, De Trez C, Vermeersch M, Perez-Morga D, Van Den Abbeele J. The dermis as a delivery dite of Trypanosoma brucei for tsetse flies. PLoS Pathog. 2016;12: e1005744. pmid:27441553
  27. 27. Hertz-Fowler C, Figueiredo LM, Quail MA, Becker M, Jackson A, Bason N, et al. Telomeric expression sites are highly conserved in Trypanosoma brucei. PLoS One. 2008;3: e3527. pmid:18953401
  28. 28. Bangs JD. Evolution of antigenic variation in African trypanosomes: Variant surface glycoprotein expression, structure, and function. Bioessays. 2018; e1800181. pmid:30370931
  29. 29. Turner CM, Barry JD, Maudlin I, Vickerman K. An estimate of the size of the metacyclic variable antigen repertoire of Trypanosoma brucei rhodesiense. Parasitology. 1988;97: 269–76. pmid:3059263
  30. 30. Lenardo MJ, Rice-Ficht AC, Kelly G, Esser KM, Donelson JE. Characterization of the genes specifying two metacyclic variable antigen types in Trypanosoma brucei rhodesiense. Proc Natl Acad Sci U S A. 1984;81: 6642–6. pmid:6593722
  31. 31. Jackson DG, Owen MJ, Voorheis HP. A new method for the rapid purification of both the membrane-bound and released forms of the variant surface glycoprotein from Trypanosoma brucei. Biochem J. 1985;230: 195–202. pmid:4052037
  32. 32. Van Den Abbeele J, Caljon G, De Ridder K, De Baetselier P, Coosemans M. Trypanosoma brucei modifies the tsetse salivary composition, altering the fly feeding behavior that favors parasite transmission. PLoS Pathog. 2010;6: e1000926. pmid:20532213
  33. 33. Alves-Silva J, Ribeiro JM, Van Den Abbeele J, Attardo G, Hao Z, Haines LR, et al. An insight into the sialome of Glossina morsitans morsitans. BMC Genomics. 2010;11: 213. pmid:20353571
  34. 34. Kariithi HM, Boeren S, Murungi EK, Vlak JM, Abd-Alla AM. A proteomics approach reveals molecular manipulators of distinct cellular processes in the salivary glands of Glossina m. morsitans in response to Trypanosoma b. brucei infections. Parasit Vectors. 2016;9: 424. pmid:27485005
  35. 35. Matetovici I, Caljon G, Van Den Abbeele J. Tsetse fly tolerance to T. brucei infection: transcriptome analysis of trypanosome-associated changes in the tsetse fly salivary gland. BMC Genomics. 2016;17: 971. pmid:27884110
  36. 36. Savage AF, Kolev NG, Franklin JB, Vigneron A, Aksoy S, Tschudi C. Transcriptome profiling of Trypanosoma brucei development in the tsetse fly vector Glossina morsitans. PLoS One. 2016;11: e0168877. pmid:28002435
  37. 37. Savage AF, Cerqueira GC, Regmi S, Wu Y, El Sayed NM, Aksoy S. Transcript expression analysis of putative Trypanosoma brucei GPI-anchored surface proteins during development in the tsetse and mammalian hosts. PLoS Negl Trop Dis. 2012;6: e1708. pmid:22724039
  38. 38. Vigneron A, O’Neill MB, Weiss BL, Savage AF, Campbell OC, Kamhawi S, Valenzuela JG, Aksoy S. Single-cell RNA sequencing of Trypanosoma brucei from tsetse salivary glands unveils metacyclogenesis and identifies potential transmission blocking antigens. Proc Natl Acad Sci USA. 2020;117: 2613–2621. pmid:31964820
  39. 39. Cheng Q, Aksoy S. Tissue tropism, transmission and expression of foreign genes in vivo in midgut symbionts of tsetse flies. Insect Mol Biol. 1999;8: 125–32. pmid:9927181
  40. 40. Jackson AP, Allison HC, Barry JD, Field MC, Hertz-Fowler C, Berriman M. A cell-surface phylome for African trypanosomes. PLoS Negl Trop Dis. 2013;7: e2121. pmid:23556014
  41. 41. Guindon S, Gascuel O. A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst Biol. 2003;52: 696–704. pmid:14530136
  42. 42. Zamze SE, Ferguson MA, Collins R, Dwek RA, Rademacher TW. Characterization of the cross-reacting determinant (CRD) of the glycosyl-phosphatidylinositol membrane anchor of Trypanosoma brucei variant surface glycoprotein. Eur J Biochem. 1988;176: 527–34. pmid:2458923
  43. 43. Webb B, Sali A. Protein structure modelling with MODELLER. Methods Mol Biol. 2021;2199: 239–255. pmid:33125654
  44. 44. Holm L, Park J. DaliLite workbench for protein structure comparison. Bioinformatics. 2000;16: 566–7. pmid:10980157
  45. 45. Higgins MK, Tkachenko O, Brown A, Reed J, Raper J, Carrington M. Structure of the trypanosome haptoglobin-hemoglobin receptor and implications for nutrient uptake and innate immunity. Proc Natl Acad Sci U S A. 2013;110: 1905–10. pmid:23319650
  46. 46. Stødkilde K, Torvund-Jensen M, Moestrup SK, Andersen CB. Structural basis for trypanosomal haem acquisition and susceptibility to the host innate immune system. Nature Comms. 2014;5: 5487. pmid:25410714
  47. 47. Freymann D, Down J, Carrington M, Roditi I, Turner M, Wiley D. 2.9 Å resolution structure of the N-terminal domain of a variant surface glycoprotein from Trypanosoma brucei. J Mol Biol. 1990;216: 141–60. pmid:2231728
  48. 48. Loveless BC, Mason JW, Sakurai T, Inoue N, Razavi M, Pearson TW, et al. Structural characterization and epitope mapping of the glutamic acid/alanine-rich protein from Trypanosoma congolense: defining assembly on the parasite cell surface. J Biol Chem. 2011;286: 20658–65. pmid:21471223
  49. 49. Ashkenazy H, Abadi S, Martz E, Chay O, Mayrose I, Pupko T, et al. ConSurf 2016: an improved methodology to estimate and visualize evolutionary conservation in macromolecules. Nucleic Acids Res. 2016;44: W344–50. pmid:27166375
  50. 50. Ashkenazy H, Erez E, Martz E, Pupko T, Ben-Tal N. ConSurf 2010: calculating evolutionary conservation in sequence and structure of proteins and nucleic acids. Nucleic Acids Res. 2010;38: W529–33. pmid:20478830
  51. 51. Roy A, Kucukural A, Zhang Y. I-TASSER: a unified platform for automated protein structure and function prediction. Nat Protoc. 2010;5: 725–38. pmid:20360767
  52. 52. Yang J, Yan R, Roy A, Xu D, Poisson J, Zhang Y. The I-TASSER Suite: protein structure and function prediction. Nat Methods. 2015;12: 7–8. pmid:25549265
  53. 53. Zhang Y. I-TASSER server for protein 3D structure prediction. BMC Bioinformatics. 2008;9: 40. pmid:18215316
  54. 54. McGuffin LJ, Atkins JD, Salehe BR, Shuid AN, Roche DB. IntFOLD: an integrated server for modelling protein structures and functions from amino acid sequences. Nucleic Acids Res. 2015;43: W169–73. pmid:25820431
  55. 55. MacGregor P, Rojas F, Dean S, Matthews KR. Stable transformation of pleomorphic bloodstream form Trypanosoma brucei. Mol Biochem Parasitol. 2013;190: 60–2. pmid:23835071
  56. 56. Rogers ME. The role of Leishmania proteophosphoglycans in sand fly transmission and infection of the mammalian host. Front Microbiol. 2012;3: 223. pmid:22754550
  57. 57. Atayde VD, Aslan H, Townsend S, Hassani K, Kamhawi S, Olivier M. Exosome secretion by the parasitic protozoan Leishmania within the sand fly midgut. Cell Rep. 2015;13: 957–67. pmid:26565909
  58. 58. Giraud E, Lestinova T, Derrick T, Martin O, Dillon RJ, Volf P, et al. Leishmania proteophosphoglycans regurgitated from infected sand flies accelerate dermal wound repair and exacerbate leishmaniasis via insulin-like growth factor 1-dependent signalling. PLoS Pathog. 2018;14: e1006794. pmid:29352310
  59. 59. Colombo M, Raposo G, Thery C. Biogenesis, secretion, and intercellular interactions of exosomes and other extracellular vesicles. Ann Rev Cell Dev Biol. 2014;30: 255–89. pmid:25288114
  60. 60. Szempruch AJ, Sykes SE, Kieft R, Dennison L, Becker AC, Gartrell A, et al. Extracellular vesicles from Trypanosoma brucei mediate virulence factor transfer and cause host anemia. Cell. 2016;164: 246–57. pmid:26771494
  61. 61. Beecroft RP, Roditi I, Pearson TW. Identification and characterization of an acidic major surface glycoprotein from procyclic stage Trypanosoma congolense. Mol Biochem Parasitol. 1993;61: 285–94. pmid:7903427
  62. 62. Bayne RA, Kilbride EA, Lainson FA, Tetley L, Barry JD. A major surface antigen of procyclic stage Trypanosoma congolense. Mol Biochem Parasitol. 1993;61: 295–310. pmid:8264732
  63. 63. Sakurai T, Sugimoto C, Inoue N. Identification and molecular characterization of a novel stage-specific surface protein of Trypanosoma congolense epimastigotes. Mol Biochem Parasitol. 2008;161: 1–11. pmid:18571746
  64. 64. Telleria EL, Benoit JB, Zhao X, Savage AF, Regmi S, Alves e Silva TL, et al. Insights into the trypanosome-host interactions revealed through transcriptomic analysis of parasitized tsetse fly salivary glands. PLoS Negl Trop Dis. 2014;8: e2649. pmid:24763140
  65. 65. Hutchinson S, Foulon S, Crouzols A, Menafra R, Rotureau B, Griffiths AD, Bastin P. The establishment of variant surface glycoprotein monoallelic expression revealed by single-cell RNA-seq of Trypanosoma brucei in the tsetse fly salivary glands. PLoS Pathog. 2021;17: e1009904. pmid:34543350
  66. 66. Doleželová E, Kunzová M, Dejung M, Levin M, Panicucci B, Regnault C, Janzen CJ, Barrett MP, Butter F, Zíková A. Cell-based and multi-omics profiling reveals dynamic metabolic repurposing of mitochondria to drive developmental progression of Trypanosoma brucei. PLoS Biol. 2020;18: e3000741. pmid:32520929
  67. 67. Lane-Serff H, MacGregor P, Lowe ED, Carrington M, Higgins MK. Structural basis for ligand and innate immunity factor uptake by the trypanosome haptoglobin-haemoglobin receptor. eLife. 2014;3: e05553. pmid:25497229
  68. 68. Graham S, Matthews K, Shiels P, Barry J. Distinct, developmental stage-specific activation mechanisms of trypanosome VSG genes. Parasitology. 1990;101: 361–7. pmid:2092293
  69. 69. McCann JR, Sheets AJ, Grass S, St Geme JW 3rd. The Haemophilus cryptic genospecies Cha adhesin has at least two variants that differ in host cell binding, bacterial aggregation, and biofilm formation properties. J Bacteriol. 2014;196: 1780–8. pmid:24584499
  70. 70. Song J, Tan H, Perry AJ, Akutsu T, Webb GI, Whisstock JC, et al. PROSPER: an integrated feature-based tool for predicting protease substrate cleavage sites. PLoS ONE 2012; 7: e50300. pmid:23209700
  71. 71. Vickerman K. On the surface coat and flagellar adhesion in trypanosomes. J Cell Sci. 1969;5: 163–93. pmid:5353653
  72. 72. Tiengwe C, Bush PJ, Bangs JD. Controlling transferrin receptor trafficking with GPI-valence in bloodstream stage African trypanosomes. PLoS Pathog. 2017;13: e1006366. pmid:28459879
  73. 73. Christiano R, Kolev NG, Shi H, Ullu E, Walther TC, Tschudi C. The proteome and transcriptome of the infectious metacyclic form of Trypanosoma brucei define quiescent cells primed for mammalian invasion. Mol Microbiol. 2017;106: 74–92. pmid:28742275
  74. 74. Pedram M, Donelson JE. The anatomy and transcription of a monocistronic expression site for a metacyclic variant surface glycoprotein gene in Trypanosoma brucei. J Biol Chem. 1999;274: 16876–83. pmid:10358033
  75. 75. Alarcon CM, Son HJ, Hall T, Donelson JE. A monocistronic transcript for a trypanosome variant surface glycoprotein. Mol Cell Biol. 1994;14: 5579–91. pmid:8035832
  76. 76. Paindavoine P, Pays E, Laurent M, Geltmeyer Y, Le Ray D, Mehlitz D, et al. The use of DNA hybridization and numerical taxonomy in determining relationships between Trypanosoma brucei stocks and subspecies. Parasitology. 1986;92: 31–50. pmid:3960593
  77. 77. Cross GA, Kim HS, Wickstead B. Capturing the variant surface glycoprotein repertoire (the VSGnome) of Trypanosoma brucei Lister 427. Mol Biochem Parasitol. 2014;195: 59–73. pmid:24992042
  78. 78. Gordon DM, McGovern TW, Krzych U, Cohen JC, Schneider I, LaChance R, et al. Safety, immunogenicity, and efficacy of a recombinantly produced Plasmodium falciparum circumsporozoite protein-hepatitis B surface antigen subunit vaccine. J Infect Dis. 1995;171: 1576–85. pmid:7769295
  79. 79. Engstler M, Boshart M. Cold shock and regulation of surface protein trafficking convey sensitization to inducers of stage differentiation in Trypanosoma brucei. Genes Dev. 2004;18: 2798–811. pmid:15545633
  80. 80. Rojas F, Silvester E, Young J, Milne R, Tettey M, Houston DR, et al. Oligopeptide signalling through TbGPR89 drives trypanosome quorum sensing. Cell. 2019;176: 306–17 e16. pmid:30503212
  81. 81. Belda E, Moya A, Bentley S, Silva FJ. Mobile genetic element proliferation and gene inactivation impact over the genome structure and metabolic capabilities of Sodalis glossinidius, the secondary endosymbiont of tsetse flies. BMC Genomics. 2010;11: 449. pmid:20649993
  82. 82. Aslett M, Aurrecoechea C, Berriman M, Brestelli J, Brunk BP, Carrington M, et al. TriTrypDB: a functional genomic resource for the trypanosomatidae. Nucleic Acids Res. 2010;38: D457–62. pmid:19843604
  83. 83. Apweiler R, Attwood TK, Bairoch A, Bateman A, Birney E, Biswas M, et al. The InterPro database, an integrated documentation resource for protein families, domains and functional sites. Nucleic Acids Res. 2001;29: 37–40. pmid:11125043
  84. 84. Lawson D, Arensburger P, Atkinson P, Besansky NJ, Bruggner RV, Butler R, et al. VectorBase: a home for invertebrate vectors of human pathogens. Nucleic Acids Res. 2007;35: D503–5. pmid:17145709
  85. 85. Sievers F, Wilm A, Dineen D, Gibson TJ, Karplus K, Li W, et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol Syst Biol. 2011;7: 539. pmid:21988835
  86. 86. Waterhouse AM, Procter JB, Martin DM, Clamp M, Barton GJ. Jalview Version 2—a multiple sequence alignment editor and analysis workbench. Bioinformatics. 2009;25: 1189–91. pmid:19151095
  87. 87. Petersen TN, Brunak S, von Heijne G, Nielsen H. SignalP 4.0: discriminating signal peptides from transmembrane regions. Nat Methods. 2011;8: 785–6. pmid:21959131
  88. 88. Pierleoni A, Martelli PL, Casadio R. PredGPI: a GPI-anchor predictor. BMC Bioinformatics. 2008;9: 392. pmid:18811934
  89. 89. Sonnhammer EL, von Heijne G, Krogh A. A hidden Markov model for predicting transmembrane helices in protein sequences. Proc Int Conf Intell Syst Mol Biol. 1998;6: 175–82. pmid:9783223.
  90. 90. Krogh A, Larsson B, von Heijne G, Sonnhammer EL. Predicting transmembrane protein topology with a hidden Markov model: application to complete genomes. J Mol Biol. 2001;305: 567–80. pmid:11152613
  91. 91. Holm L, Rosenstrom P. Dali server: conservation mapping in 3D. Nucleic Acids Res. 2010;38: W545–9. pmid:20457744
  92. 92. Brenndorfer M, Boshart M. Selection of reference genes for mRNA quantification in Trypanosoma brucei. Mol Biochem Parasitol. 2010;172: 52–5. pmid:20302889
  93. 93. Dewar S, Sienkiewicz N, Ong HB, Wall RJ, Horn D, Fairlamb AH. The role of folate transport in antifolate drug action in Trypanosoma brucei. J Biol Chem. 2016;291: 24768–78. pmid:27703008
  94. 94. Kelly S, Reed J, Kramer S, Ellis L, Webb H, Sunter J, et al. Functional genomics in Trypanosoma brucei: a collection of vectors for the expression of tagged proteins from endogenous and ectopic gene loci. Mol Biochem Parasitol. 2007;154: 103–9. pmid:17512617
  95. 95. Burkard G, Fragoso CM, Roditi I. Highly efficient stable transformation of bloodstream forms of Trypanosoma brucei. Mol Biochem Parasitol. 2007;153: 220–3. pmid:17408766
  96. 96. Beneke T, Madden R, Makin L, Valli J, Sunter J, Gluenz E. A CRISPR Cas9 high-throughput genome editing toolkit for kinetoplastids. R Soc Open Sci. 2017;4: 170095. pmid:28573017
  97. 97. Leslie AGW. Recent changes to the MOSFLM package for processing film and image plate data. Joint CCP4 + ESF-EAMCB Newsletter on Protein Crystallography. 1992;26.
  98. 98. Evans PR. Scaling and assessment of data quality. Acta Cryst. 2005;D: 72–82. pmid:16369096
  99. 99. Collaborative Computational Project N4. The CCP4 suite: programs for protein crystallography. Acta Crystallogr Sect D Biol Crystallogr. 1994;50: 760–3. pmid:15299374
  100. 100. McCoy AJ, Grosse-Kunstleve RW, Adams PD, Winn MD, Storoni LC, Read RJ. Phaser crystallographic software. J Appl Crystallogr. 2007;40: 658–74. pmid:19461840
  101. 101. Emsley P, Cowtan K. Coot: model-building tools for molecular graphics. Acta Crystallog sect D. 2004;60: 2126–32. pmid:15572765
  102. 102. Adams PD, Afonine PV, Bunkoczi G, Chen VB, Davis IW, Echols N, et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr D Biol Crystallogr. 2010;66: 213–21. pmid:20124702
  103. 103. Lovell SC, Davis IW, Arendall WB 3rd, de Bakker PI, Word JM, Prisant MG, et al. Structure validation by Calpha geometry: phi,psi and Cbeta deviation. Proteins. 2003;50: 437–50. pmid:12557186
  104. 104. Benkert P, Tosatto SC, Schomburg D. QMEAN: A comprehensive scoring function for model quality assessment. Proteins. 2008;71: 261–77. pmid:17932912
  105. 105. Wiederstein M, Sippl MJ. ProSA-web: interactive web service for the recognition of errors in three-dimensional structures of proteins. Nucleic Acids Res. 2007;35: W407–10. pmid:17517781
  106. 106. Rose C, Belmonte R, Armstrong SD, Molyneux G, Haines LR, Lehane MJ, et al. An investigation into the protein composition of the teneral Glossina morsitans morsitans peritrophic matrix. PLoS Negl Trop Dis. 2014;8: e2691. pmid:24763256
  107. 107. Matyas GR, Muderhwa JM, Alving CR. Oil-in-water liposomal emulsions for vaccine delivery. Methods Enzymol. 2003;373: 34–50. pmid:14714395
  108. 108. Portillo S, Zepeda BG, Iniguez E, Olivas JJ, Karimi NH, Moreira OC, et al. A prophylactic alpha-Gal-based glycovaccine effectively protects against murine acute Chagas disease. NPJ Vaccines. 2019;4: 13. pmid:30911415