Skip to main content

Changes in saliva protein profile throughout Rhipicephalus microplus blood feeding

Abstract

Background

When feeding on a vertebrate host, ticks secrete saliva, which is a complex mixture of proteins, lipids, and other molecules. Tick saliva assists the vector in modulating host hemostasis, immunity, and tissue repair mechanisms. While helping the vector to feed, its saliva modifies the site where pathogens are inoculated and often facilitates the infection process. The objective of this study is to uncover the variation in protein composition of Rhipicephalus microplus saliva during blood feeding.

Methods

Ticks were fed on calves, and adult females were collected, weighed, and divided in nine weight groups, representing the slow and rapid feeding phases of blood feeding. Tick saliva was collected, and mass spectrometry analyses were used to identify differentially secreted proteins. Bioinformatic tools were employed to predict the structural and functional features of the salivary proteins. Reciprocal best hit analyses were used to identify conserved families of salivary proteins secreted by other tick species.

Results

Changes in the protein secretion profiles of R. microplus adult female saliva during the blood feeding were observed, characterizing the phenomenon known as “sialome switching.” This observation validates the idea that the switch in protein expression may serve as a mechanism for evading host responses against tick feeding. Cattle tick saliva is predominantly rich in heme-binding proteins, secreted conserved proteins, lipocalins, and protease inhibitors, many of which are conserved and present in the saliva of other tick species. Additionally, another remarkable observation was the identification of host-derived proteins as a component of tick saliva.

Conclusions

Overall, this study brings new insights to understanding the dynamics of the proteomic profile of tick saliva, which is an important component of tick feeding biology. The results presented here, along with the disclosed sequences, contribute to our understanding of tick feeding biology and might aid in the identification of new targets for the development of novel anti-tick methods.

Graphical Abstract

Background

Ticks are pool feeders that accomplish feeding by lacerating small blood vessels and ingesting the blood that flows from the feeding site. The hard tick feeding cycle lasts a few days and includes three phases: (i) the preparatory feeding phase, during which the tick attaches onto host's skin and creates the feeding lesion; (ii) the slow feeding phase, when the tick ingests moderate amounts of blood, begins to transmit pathogens, and grows new tissue to prepare itself for (iii) the rapid feeding phase, when the tick feeds until it reaches repletion [1]. This feeding style triggers several host-derived responses, and ticks face the problem of wound healing, which is achieved through four precisely and highly programmed phases: hemostasis, inflammation, cell proliferation and migration, and tissue remodeling. Together, these phases are barriers to acquiring a blood meal [2, 3]. To overcome these barriers, tick salivary glands have evolved a complex and sophisticated pharmacological armamentarium consisting of bioactive molecules to assist blood feeding [4].

While assisting the vector in feeding, tick saliva also modifies the site where pathogens are injected and, in many cases, facilitates the infection process [5]. Tick saliva contains a mixture of secretions from different salivary gland acini, making its composition variable and influenced by various factors [2]. It is evident that the kinetics and amount of saliva secretion differ considerably during the feeding process [6,7,8,9]. This effect is attributed to host and environmental stimuli, morphological and physiological alterations of salivary glands during the parasitic period, among other factors [2].

Transcriptomic and proteomic analyses of ticks have enabled the study of gene expression changes in different tick species and have been instrumental in understanding tick physiology and the biology of salivary glands and other tissues [10, 11]. Bioinformatic analyses of sequencing data are used to identify proteins and predict their function, providing a better understanding of how blood feeding and pathogen infection impact tick physiology. Although some progress has been made in recent years [11, 12], information on how tick saliva composition is modified during the tick feeding process is still scarce. For the cattle tick Rhipicephalus microplus, which is a one-host tick that feeds on bovines and is considered one of the most harmful cattle parasites in subtropical areas worldwide [13], our group pioneered the characterization of its saliva proteome by comparing saliva from partially and fully engorged females [9]. In the present study, we conducted a comprehensive analysis to evaluate changes in protein content profiles of R. microplus saliva in adult females throughout the entire feeding process. Uniquely, instead of grouping the different feeding stages based on partially and fully engorged ticks, we divided the ticks into nine groups based on their weight. Dividing ticks into groups by weight provides a better indicator of the physiological status of the tick feeding process [14]. The analyzed groups in this study represent the slow and rapid feeding phases of blood feeding. Mass spectrometry was used to identify differentially secreted proteins in tick saliva, and bioinformatic tools were employed to predict the structural and functional features of the salivary proteins. This work provides new insights into the plasticity of salivary protein secretion and the identification of compounds that may play a role in tick biology. Furthermore, this catalog of secreted salivary proteins serves as the basis for subsequent studies conducted by many other research groups.

Methods

Ticks and saliva collection

Rhipicephalus microplus ticks (Porto Alegre strain), free of pathogens such as Babesia spp. and Anaplasma spp., were reared on Hereford calves (Bos taurus taurus) that were brought from a naturally tick-free area (Santa Vitória do Palmar, RS, Brazil; 33°32′2"S, 53°20′59"W) and maintained in individual boxes. Calves were infested with approximately 20,000 10-day-old larvae (from 1 g of R. microplus eggs), and after 21 days, adult females attached to the hosts were manually collected [15]. Adult females were collected, weighted, and grouped in nine groups (weight ± SD): group Rm-1 (34 ticks, weight 6.8 ± 1.1 mg); group Rm-2 (22 ticks, weight 15.8 ± 1.3 mg); group Rm-3 (16 ticks, weight 25.1 ± 1.3 mg); group Rm-4 (40 ticks, weight 35.0 ± 2.5 mg); group Rm-5 (25 ticks, weight 52.5 ± 2.8 mg); group Rm-6 (22 ticks, weight 88.0 ± 12.6 mg); group Rm-7 (16 ticks, weight 172.5 ± 16.1 mg); group Rm-8 (39 ticks, weight 270.2 ± 14.9 mg); and a group representing fully engorged females, group FEF (12 ticks, weight 347.5 ± 9.0 mg).

Ticks were rinsed with sterile water, dried on a paper towel, and placed dorsal side down on a glass slide containing tape. Salivation was induced by injecting 1–5 μl of 2% pilocarpine hydrochloride (in phosphate-buffered saline, pH 7.4) on the ventral side of the lower right coxa using a Hamilton syringe (Hamilton Co., Reno, NV, USA) [16]. Subsequently, after the injections, saliva was periodically collected using a Hamilton syringe (every 15 min over approximately 4 h). Protein concentration was determined using the bicinchoninic acid method (BCA Protein Assay, Pierce, Rockford, IL, USA) and stored at − 70 °C for later use.

Protein digestion and sample preparation

The saliva of R. microplus ticks from each specific feeding group was digested in a solution with trypsin. First, the saliva was diluted in 8 M urea/0.1 M Tris, pH 8.5, reduced with 5 mM Tris(2-carboxyethyl) phosphine hydrochloride (TCEP, Sigma-Aldrich, St Louis, MO, USA) and alkylated with 25 mM iodoacetamide (Sigma-Aldrich). Proteins were digested overnight at 37 °C in 2 M urea/0.1 M Tris pH 8.5, 1 mM CaCl2 with trypsin (Promega) at a final ratio of 1:20 (enzyme:substrate). Digestion reactions, at a final concentration of 0.15 µg/ml, were quenched with formic acid (5% final concentration) and centrifuged to remove debris.

Pre-columns and analytical columns

Reversed phase pre-columns were prepared by first creating a Kasil frit at one end of a deactivated 250 µm ID/360 µm OD capillary (Agilent Technologies, Santa Clara, CA, USA). Kasil frits were prepared by dipping 20 cm capillary in 300 µl Kasil 1624 (PQ Corporation, Malvern, PA, USA) and 100 µl formamide solution, curing at 100 °C for 3 h, and adjusting the length. Pre-columns were packed in house (John Yates III's laboratory at the Scripps Research Institute) with 2 cm of 5 µm ODS-AQ C18 (YMC America, INC., Allentown, PA, USA) particles from particle slurries in methanol. Analytical reversed-phase columns were fabricated by pulling a 100-µm ID/360 µm OD silica capillary (Molex Polymicro Technologies™, Austin, TX, USA) to a 5-µm ID tip. The same packing material was packed for a length of 20 cm directly behind the pulled tip. Reversed-phase pre-columns and analytical columns were connected using a zero-dead volume union (IDEX Corp., Upchurch Scientific, Oak Harbor, WA, USA).

LC-MS/MS

Peptide mixtures were analyzed by nanoflow liquid chromatography mass spectrometry using an Easy NanoLC II and a Q Exactive mass spectrometer (Thermo Scientific, Waltham, MA, USA). Peptides eluted from the analytical column were electrosprayed directly into the mass spectrometer. Buffer A and B consisted of 5% acetonitrile/0.1% formic acid and 80% acetonitrile/0.1% formic acid, respectively. The flow rate was set to 400 nl/min. Saliva samples (1.5 µg per injection) were separated in 155-min chromatographic runs, as follows: a 1–10% gradient of buffer B in 10 min, a 10–40% gradient of buffer B in 100 min, a 40–50% gradient of buffer B in 10 min, and a 50–90% gradient of buffer B in 10 min. The column was held at 90% buffer B for 10 min, reduced to 1% buffer B, and re-equilibrated prior to the next injection.

The mass spectrometer was operated in a data-dependent mode, collecting a full MS scan from 400 to 1200 m/z at 70,000 resolution and an AGC target of 1 × 106. The 10 most abundant ions per scan were selected for MS/MS at 17,500 resolution and AGC target of 2 × 105 and an underfill ratio of 0.1%. Maximum fill times were 20 and 120 ms for MS and MS/MS scans, respectively, with dynamic exclusion of 15 s. Normalized collision energy was set to 25. Three technical replicates were used to analyze the protein extracts from tick saliva samples.

Data analysis

Tandem mass spectra were extracted from Thermo RAW files using RawExtract 1.9.9.2 [17] and searched with ProLuCID [18] against a non-redundant database. The database consisted of an R. microplus database (22,009 entries) [19] concatenated with a B. taurus Uniprot reference database (23,868 entries) and reverse sequences of all entries. The searches were performed using the Integrated Proteomics Pipeline—IP2 (Integrated Proteomics Applications, Inc.). The search space included all fully tryptic and half-tryptic peptide candidates. Carbamidomethylation of cysteine was used as a static modification. The data were searched with a 50-ppm precursor ion tolerance and a 20-ppm fragment ion tolerance.

The validity of the peptide spectrum matches (PSMs) generated by ProLuCID was assessed using the Search Engine Processor (SEPro) module from the PatternLab for Proteomics platform [20]. Identifications were grouped by charge state and tryptic status, resulting in four distinct subgroups. For each group, ProLuCID XCorr, DeltaCN, DeltaMass, ZScore, the number of peaks matched, and secondary rank values were used to generate a Bayesian discriminating function. A cutoff score was established to accept a false discovery rate (FDR) of 1% based on the number of decoys. This procedure was independently performed on each data subset, resulting in a false-positive rate that was independent of tryptic status or charge state. Additionally, a minimum sequence length of six residues per peptide was required. The results were post-processed to only accept PSMs with < 10 ppm precursor mass error. Furthermore, only proteins identified in two out of the three technical replicates were considered for functional annotation.

Protein functional annotation and classification

For annotation, we used an in-house program that scans a vocabulary of approximately 400 words and their order of appearance in the protein matches from BLASTp results, including their similarities and coverage [21]. Automated annotation of the tick proteins was based on matches to various databases, including Chelicerata from NCBI (https://www.ncbi.nlm.nih.gov/genbank/) and UniProt (https://www.ebi.ac.uk/uniprot/), UniProtKB (https://www.uniprot.org/help/uniprotkb), CDD, COG, KOG, PFAM, PRK, TIGR, SMART (ftp://ftp.ncbi.nih.gov/pub/mmdb/cdd/little_endian/), MEROPS (https://www.ebi.ac.uk/merops/), RefSeq invertebrate (https://ftp.ncbi.nlm.nih.gov/refseq/release/invertebrate/), and the TickSialoFam [11]. A database containing markers for extracellular vesicles was also used [22]. Signal peptide, furin cleavage, transmembrane, GPI anchors, and glycosylation site predictions were determined using software from the Center for Biological Sequence Analysis (https://www.cbs.dtu.dk/services/).

The identity to other tick proteins was evaluated by BLASTp using coding sequences extracted from tick genomes including Dermacentor silvarum, Haemaphysalis longicornis, Hyalomma asiaticum, Ixodes persulcatus, I. scapularis, Rhipicephalus microplus, and R. sanguineus [23].

Graphical visualization

Normalized spectral abundance factors (NSAF) were used to represent relative abundance and secretion dynamics. Secretion dynamic plots were generated using the ggplot2 library [24]. To verify whether the proteins identified in the different groups were organized in clusters, the NSAF values of the tick proteins were submitted to the CLICK algorithm of the Expander program [25]. For heatmaps, NSAF values were normalized by calculating the Z-score, and these values were used to generate heatmaps using the heatmap2 function from the ggplot2 library. Principal component analysis (PCA) was performed using NSAF values and the prcomp function from the ggfortify library [26].

Structure prediction of salivary proteins

The Alphafold2 program [27] was used to predict the tertiary structures of R. microplus salivary proteins that were identified. The program was run locally on the NIH Biowulf cluster in a Linux environment using the monomer mode. The Dali program [28] was used to compare the Alphafold2 predictions to the structures available in the PDB database. The program was run locally in the NIH Biowulf cluster. Structures were visualized, and figures were generated using PyMol (PyMOL Molecular Graphics System, version 2.6.0a0, Schrödinger, LLC). The spreadsheet (Additional file 3: Table S3—sheet tab “AlphaFold”) has links to pdb files, which need programs that can open them. We suggest the use of ChimeraX, available online (https://www.cgl.ucsf.edu/chimerax/download.html).

Identification of reciprocal best hits in other tick saliva proteomes

To identify reciprocal best hits with tick saliva proteins described in other studies, a BLASTp analysis was performed using pairwise analysis. A final cutoff value of 1e-6 and a coverage of at least 50% were selected [29]. Databases were retrieved from studies describing proteomic studies of saliva from Ornithodoros moubata males and females [30], R. microplus partially and fully engorged females [9], H. longicornis nymphs and adult females [8], I. scapularis and Amblyomma americanum adult females [7, 31], I. scapularis and A. americanum unfed adult females exposed to different vertebrate hosts [32], Amblyomma sculptum adult females [33], and I. scapularis nymphs infected and non-infected with Borrelia burgdorferi [34] and from cement from A. americanum [6] and I. scapularis [35].

Alignment and sequence analysis

Protein sequence alignments were performed using Clustal W within the BioEdit 7.2.6.1 [36] and visualized with GeneDoc version 2.7 [37].

Results and discussion

An overview on Rhipicephalus microplus saliva proteome throughout the blood feeding

The R. microplus feeding cycle spans several days and consists of three distinct phases [1, 38]. The present study complements a pioneering study developed by our group when we compared the proteomic profile of saliva from partially and fully engorged R. microplus females [9]. Ticks were fed on calves, and adult females were collected, weighed, and divided into nine groups. Uniquely, instead of grouping the different feeding stages by days of feeding, we grouped the ticks by their weight, as it is a better indicator of the physiological status of the tick feeding process [19, 39]. Ticks in these engorgement stages represent feeding phases that adult ticks pass through during the blood-feeding process, including representatives from the slow feeding phase, rapid feeding phase, and fully engorged ticks (Fig. 1).

Fig. 1
figure 1

Group of Rhipicephalus microplus ticks collected through blood feeding. A Adult R. microplus females were collected, weighed, and divided into nine groups as follows (mean ± SD): (Rm-1) 6.8 ± 1.1 mg; (Rm-2) 15.8 ± 1.3 mg; (Rm-3) 25.1 ± 1.3 mg; (Rm-4) 35.0 ± 2.5 mg; (Rm-5) 52.5 ± 2.8 mg; (Rm-6) 88.0 ± 12.6 mg; (Rm-7) 172.5 ± 16.1 mg; (Rm-8) 270.2 ± 14.9 mg; and a group representing fully engorged females (FEF) weighing 347.5 ± 9.0 mg. A graphical representation demonstrates the grouping of ticks according to weight and their relationship with the feeding period. This graph illustrates the initial slow-feeding phase (from Rm-1 to Rm-6), followed by the rapid-feeding phase (from Rm-6 to FEF). B From left to right, a series of images showing a representative tick from each group, spanning from Rm-1 to FEF

Pilocarpine-induced saliva was successfully harvested from these ticks, and protein identification was performed by LC–MS/MS. The generated raw data from the shotgun proteomic approach were used to query the R. microplus de novo transcriptome assembly using PatternLab software, which allows data normalization by the normalized spectral abundance factor (NSAF) approach. The search of extracted tandem mass spectra against the tick and host protein databases produced hits to 346 tick-derived proteins (Additional file 1: Table S1) and 74 host-derived proteins (Additional file 2: Table S2). When subjected to validity analysis in the PatternLab for Proteomics platform, 284 of the 346 tick-derived proteins were considered true proteins since they were detected in a minimum of two out of the three runs (Additional file 1: Table S1—sheet tab “(Additional file 1: Table S1A”), while the remaining 62 proteins detected in only one of the three runs were considered low confidence and were not further analyzed (Additional file 1: Table S1—sheet tab “Additional file 1: Table S1B”). Of the 74 host-derived proteins detected in R. microplus saliva, 47 met the criteria for authentication (Additional file 2: Table S2—sheet tab “Table S2A”), and the remaining 27 were not further analyzed (Additional file 2: Table S2—sheet tab “Table S2B”). The 284 tick-derived proteins were annotated and classified into 22 different functional classes (Additional file 3: Table S3).

Insight into the sialome switch of Rhipicephalus microplus

During the process of feeding, the salivary glands play an essential role by secreting a unique collection of molecules, forming a complex and ever-changing cocktail. This mixture exhibits a dynamic composition, with the amounts of each protein constantly varying in response to various factors, including developmental stage, blood meal uptake, time/duration of the feeding process, the host where ticks are feeding, and whether they are infected or not with different pathogens [5, 32, 40, 41], the phenomenon called "sialome switching” [42, 43].

To gain insight into the broad relationships of secretion dynamics of tick proteins with tick feeding processes, the Z-score was used to represent NSAF values, which were visualized on a heatmap. The heatmap displayed how dynamic the expression of salivary proteins during blood feeding is (Fig. 2A). Indeed, the PCA plot (Fig. 2B) showed remarkable clustering of the replicates and distinct proteomic profiles among all groups, revealing the phenomenon of "sialome switching.” The 284 tick-derived proteins were annotated and classified into 22 different functional classes. Notably, the saliva of R. microplus contained a variety of highly abundant proteins, some of which play important roles in blood feeding. These proteins include those associated with heme/iron metabolism, lipocalins, secreted conserved proteins, protease inhibitors, proteases, immunity, and the extracellular matrix (Fig. 3A). The other 15 classes are less abundant; little is known about their role in tick feeding (Fig. 3B), and they will not be discussed in detail.

Fig. 2
figure 2

An overview of Rhipicephalus microplus saliva proteome throughout blood feeding. A Heat map of R. microplus tick saliva proteome throughout blood feeding. This heat map illustrates the changes in the R. microplus tick saliva proteome throughout blood feeding. Blue indicates downregulation, while red indicates upregulation. B Principal component analysis (PCA) of R. microplus tick saliva proteome throughout blood feeding

Fig. 3
figure 3figure 3

Expression patterns of the functional classes identified in the Rhipicephalus microplus saliva proteome throughout blood feeding. Each data point represents the average NSAF value for each functional group, with error bars denoting the standard error. A Represents classes of proteins with higher abundance. B Represents classes of proteins with lower abundance

Hemelipoproteins

As previously demonstrated [9], R. microplus saliva contains a high concentration of hemelipoproteins, with a specific focus on the HeLp protein [9, 44, 45]. HeLp possesses the capability of heme binding, in addition to cholesterol, phospholipids, and free fatty acids [44, 45]. Hemelipoproteins, as classified in the heme/iron metabolism plot in Fig. 3, are one of the most abundant proteins found in the saliva of R. microplus during the slow and rapid feeding phases (Rm-1 to Rm-7). However, their presence in saliva decreases as the feeding process progresses towards the later stages of engorgement (Rm-8 and FEF). Although only six proteins belonging to this class have been identified, these six proteins contribute to most spectral counts observed in tick saliva (Additional file 1: Table S1 and Additional file 3: Table S3).

Given the primary localization of HeLp in the hemolymph of ticks [44, 45], its presence in the saliva could be attributed to the incorporation of hemolymph components by the salivary glands. However, research focusing on the transcriptional profile and protein localization of these hemelipoproteins in the salivary glands of adult and unfed ticks from various species has suggested that they may have diverse functions during blood feeding [46, 47]. Furthermore, the identification of these proteins in the saliva of other tick species demonstrates their conservation across different ticks [7, 8, 31, 48,49,50]. These findings suggest a crucial role for HeLp in tick feeding. HeLp possesses the capability to bind eight heme molecules [44, 45]. Consequently, its inoculation into the host can reduce the concentration of free heme at the feeding site, effectively preventing inflammation [51]. Additionally, this protein can bind to other molecules, including cholesterol, phospholipids, and free fatty acids [44, 45], and may be molecules yet unknow. It is plausible that these compounds could be carried at the feeding site bound to HeLp, potentially carrying small pharmacologically active molecules from salivary origin. Similarly, as suggested for the sequestration of heme at the feeding site, HeLp acting as a kratagonist of other small host-derived molecules [52] at the feeding site cannot be ruled out. Under these circumstances, hemelipoproteins may serve as heme transporters and/or kratagonists during the onset of hemoglobin digestion in the midgut. The role of these proteins in the interaction between ticks and their hosts remains largely unknown. Given their significant presence in tick saliva, conducting studies to elucidate their functions during the blood meal is crucial.

Lipocalins

Lipocalins are also highly abundant in tick saliva, with their levels increasing during the feeding process and reaching a peak in the saliva of ticks from the group Rm-5, remaining steady until the end of the feeding process (Fig. 3A). In total, R. microplus ticks secrete a minimum of 59 different lipocalins in their saliva. This observation is in accordance with studies describing the high abundance of lipocalins in the saliva of different tick species [7,8,9, 31, 48, 53].

Lipocalins are proteins composed of approximately 200 amino acids, known for their compact folding into a β-barrel shape. Within this structure, a central pocket capable of binding small hydrophobic molecules exists. In ticks, lipocalins play a crucial role as kratagonists, effectively scavenging histamine, serotonin, leukotriene B4, leukotriene C4, thromboxane A2, and cholesterol. By modulating the host's hemostasis, inflammation, and other immune responses, these lipocalins facilitate blood feeding [54]. When subjected to a scan using the Pfam database, it was revealed that, out of the 59 identified and classified as lipocalins, at least 43 possess the tick histamine-binding domain (PF02098) (Additional file 3: Table S3). Despite the known kratagonist activities of tick lipocalins, the biochemical characterization of lipocalins from R. microplus is lacking, thus hindering the identification of their primary ligands.

Proteases

A total of 25 proteases belonging to various families were identified in the saliva of R. microplus. These proteases include 12 metalloproteases, 6 serine proteases, 5 asparagine proteases, and 2 cysteine proteases. Searching the MEROPS database showed that the 12 metalloproteases can be classified into families M12B (n = 9), M17 (n = 2), and M13 (n = 1), while the asparagine and cysteine proteases are categorized into families A1 and C1, respectively. The serine proteases identified in this study were assigned to families S01 (n = 3) and S10 (n = 3) (Additional file 3: Table S3).

Notably, the saliva from ticks in the Rm-1 group contained the highest abundance of proteases. Their abundance decreased from groups Rm-2 to Rm-4, increased in group Rm-5, and remained relatively steady until the end of the feeding period (Fig. 3A). Examining the expression dynamics of proteases according to their class shows that metalloproteases, including members from M12B, M17, and M13 families, reach their peak at the start of feeding exclusively in the Rm-1 group. Their levels sharply decline in the Rm-2 group and remain consistently low until the end of the feeding period (Additional file 4: Fig. S1). The predominance of family M12B metalloproteases secreted by R. microplus at the early stages of blood feeding indicates their significance in tick feeding physiology. Tick metalloproteases have activities that promote tick feeding [55], and their blockage by silencing or immunization has an impact on tick feeding [56,57,58]

Since the discovery of the metalloproteoid family [59], we have decided to reanalyze the metalloprotease family to determine if each sequence contains the Zn-binding domain. Thus, we conducted a search and manually checked for the motif H-E-x(2)-H–x(2)-G-x(2)-H [59]. Among the 12 proteins identified within the M12B, M13, and M17 families, Rm-1513, Rm-156308, Rm-18262, Rm-24243, and Rm-2425 were found not to possess the Zn-binding domain and could be classified as part of the metalloproteoid family (Additional file 5: Fig. S2). Metalloproteases exhibit a wide range of substrate specificity, including coagulation factors, platelet membrane receptors, and von Willebrand factor [60]. It is possible that tick metalloproteoids function by specifically binding to host proteins without hydrolyzing them, but potentially inhibit their function, thereby acting as pseudoenzymes or kratagonists [59].

The proteases belonging to the S01 family exhibit a peak within the Rm-1 group, followed by a gradual reduction through the Rm-5 group. However, their levels rise again in the Rm-8 group before declining once more towards the end of the blood feeding (Additional file 4: Fig. S1). On the other hand, the proteases affiliated with the S10 family demonstrate a peak between the Rm-4 and Rm-6 groups. Subsequently, their levels decline and then remain steady until the end of the blood feeding (Additional file 4: Fig. S1). Upon examining the primary and tertiary structure of the members within the S01 family, it became apparent that Rm-77083 and Rm-163513 exhibit the characteristic folding pattern of serine proteases, featuring the catalytic triad comprising Asp, His, and Ser. However, notably, Rm-2605, despite being annotated in this family, deviates from this pattern by substituting Ser in the catalytic triad with Gly (Additional file 6: Fig. S3) [61]. The presence of trypsin-like proteases in tick saliva, especially members within the S01 family, may be a strategy to interfere with host inflammation and blood clotting similarly to what has been reported for I. scapularis saliva [62].

Proteases associated with the C01 family display a peak at the beginning of the feeding period (Rm-1 to Rm-3), which is subsequently followed by a decrease in groups Rm-4 and Rm-5. Their levels then rise again in the Rm-6 and Rm-7 groups before declining once more towards the end of the blood meal (Additional file 1: Fig. S1). On the other hand, the proteases associated with the A01 family are not expressed at the beginning of feeding; instead, their presence in the saliva gradually increases from Rm-4 to Rm-5 and remains constant until the end of feeding (Additional file 1: Fig. S1). During the examination of the primary and tertiary structure of the members within the C01 family, it became evident that both proteins within this family exhibit the characteristic folding pattern of cysteine proteases, including the catalytic dyad comprising Cys and His, as depicted in Additional file 7: Fig. S4. Most members of the C01 and A01 families have been well characterized as important players during blood meal digestion in the midgut [63, 64] and vitellin digestion during embryogenesis [65, 66]. However, their role in the saliva remains unclear. Further characterization of the tick saliva proteases identified in this study holds great promise for gaining valuable insights into their crucial role in the blood feeding process.

Protease inhibitors

Regarding protease inhibitors, in a similar fashion to the observation of proteases, a total of 26 protease inhibitors were identified in the saliva of R. microplus (Additional file 3: Table S3). However, they exhibited a higher abundance compared to the proteases (Fig. 3A). Based on the literature and information from the MEROPS database, these protease inhibitors were classified into eight families: serpins (n = 9), which include RmS-7, RmS-15, and various isoforms of RmS-3, RmS-6, and RmS-17 [9], trypsin-inhibitor like (TIL) (n = 5), Kunitz type (n = 3), cystatins (n = 3), BmSEI-like (n = 3), thyropins (n = 2), and Kazal type (n = 1).

Serpin is a class of protease inhibitors which regulates the activity of serine proteases by forming irreversible covalent complexes with them [67]. Serpins were already described in the saliva of ticks and specifically play a crucial role in tick-host interactions by modulating the host responses against tick feeding [68]. The reactive center loop (RCL) of serpins consists of a flexible stretch of 21 amino acid residues, serving as a pseudo-substrate for the target protease. The P1 residue within the RCL plays a crucial role in determining the specificity of a serpin for a particular protease [69]. Tick serpins identified in this study exhibit different P1 residues, including basic side chains (Lys for RmS-17 and Arg for RmS-6 and RmS-15), polar side chains (Ser for RmS-7), and hydrophobic side chains (Leu for RmS-3) [9], potentially suggesting these serpins target different proteases during blood feeding. The serpins identified in the saliva of R. microplus exhibit a basal content at the initiation of feeding (Rm-1 to Rm-5), which is subsequently followed by an increase in groups Rm-6 and Rm-7. However, their levels decrease again in group Rm-8 and rise once more in group FEF (Additional file 8: Fig. S5). Serpins RmS-3, RmS-6, RmS-15, and RmS-17 have been biochemically characterized and possess anticoagulant and immunomodulatory properties, suggesting the potential for these serpins to target host responses during blood feeding [70,71,72].

In arthropods, most trypsin inhibitor-like proteins (TILs) are composed of a single domain featuring two roughly perpendicular β-sheets, turns, and a long-exposed loop that encompasses the protease binding site (RCL). The protein is reinforced by five disulfide bridges, which enhance the stability of both the protein scaffold and the RCL. Furthermore, the RCL demonstrates similarities to those observed in other canonical serine protease inhibitors [73]. In this study, three TILs (Rm-10457, Rm-18594, and Rm-32449) were found to have a single domain, while two TILs (Rm-53166 and Rm-53167) exhibited two domains separated by a flexible loop. TILs with a single domain featured Leu (Rm-10457) and Ala residues (Rm-18594 and Rm-32449) at the P1 position. Interestingly, TILs with two domains have Tyr residues at the N-terminal domain and Ala residues at the C-terminal domain as their respective P1 residues (Additional file 9: Fig. S6), suggesting the potential to target two different proteases. The presence of proteins of this family in saliva has a peak at the beginning feeding (Rm-1 to Rm-3), which is then followed by a decrease in expression in groups Rm-4 to Rm-6. However, their levels rise again in group Rm-7 and continue to increase until the end of feeding (Additional file 8: Fig. S5). Only a few TIL proteins have been functionally characterized in ticks. One such protein, ixodidin, which we identified here (Rm-10457), was previously isolated from R. microplus hemocytes. It exhibits antimicrobial activity and the ability to inhibit elastase and chymotrypsin [74]. Another R. microplus protein containing a TIL domain, named BmSI-7, was also identified in our study (Rm-32449). BmSI-7 displays inhibitory activities against subtilisin A and human neutrophil elastase, suggesting a potential role in modulating inflammation at the tick feeding site on the host's skin [75]. Notably, antibodies against BmSI-7 have been shown to affect female ticks' reproductive parameters in an artificial tick feeding assay [76].

Cystatins, which act as cysteine protease inhibitors, are present in various organisms and are classified into three subfamilies based on their primary structure. In ticks, both intracellular (type 1) and secretory (type 2) cystatins have been identified and characterized. Here, three type 2 cystatins have been identified: Rmcystatin-2, BrBmcys2a, and BrBmcys2b. These cystatins inhibit host cathepsin B, cathepsin C, and cathepsin L, potentially regulating the host's immune response [77,78,79]. Similarly, members of the thyropin protein family are characterized by the presence of thyroglobulin type-1 domain repeats. A well-characterized thyropin protein was initially described in a sea anemone [80] and has been demonstrated to inhibit both cysteine and cation-dependent proteases, including cathepsin L, cathepsin S, papain, and cruzipain [80]. Despite this, the functional characterization of thyropins in ticks is currently lacking. However, the presence of these proteins in tick saliva strongly suggests their potential involvement in modulating host immune responses [81]. When analyzing the saliva profiles of these two inhibitor families, a distinct difference becomes evident. Cystatins exhibit nearly constant basal levels throughout the entire blood meal, whereas thyropins display a peak at the beginning of feeding (Rm-1 to Rm-4) followed by a gradual decline to sustained levels until the end of the meal (Additional file 8: Fig. S5).

Members of the Kunitz-type family have been extensively studied in ticks as inhibitors of numerous serine proteases, affecting blood coagulation and immune responses [82, 83]. In this study, three proteins containing Kunitz-type domains were identified. Interestingly, only one protein (Rm-51801) displayed a structure resembling the canonical Kunitz-type inhibitors, whereas the other two (Rm-4223 and Rm-770) exhibited non-canonical structures (Additional file 3: Table S3—sheet tab “AlphaFold”). Their presence in saliva peaks in the Rm-1 group drops to zero in the Rm-2 to Rm-5 groups, and then rises again in the Rm-6 group, remaining relatively constant until the end of the blood meal (Additional file 8: Fig. S5).

A single Kazal domain-containing protein was identified in the saliva, with exclusive expression observed in the Rm-1 group (Additional file 8: Fig. S5). Despite the previous classification of BmSEI as a protease inhibitor, recent studies have revealed that these proteins lack inhibitory activity but possess antimicrobial properties [76]. Notably, the structure of BmSEI (Rm-928) reveals the presence of alpha helices (Additional file 3: Table S3—sheet tab “AlphaFold”), which are characteristic of antimicrobial peptides [84].

Extracellular matrix/cell adhesion

Extracellular matrix/cell adhesion proteins encompass a large group of proteins that play roles in both structural and non-structural functions, contributing to the organization of cells and tissues [85]. In R. microplus saliva, the number of extracellular matrix proteins is elevated during the initial phases of blood feeding and decreases upon the completion of the process (Fig. 3A). Proteins within this class are represented by tick cement proteins such as glycine-rich proteins, followed by alanine- and proline-rich proteins, as well as laminin. Glycine-rich proteins have been implicated in the formation of the cement cone [35], and their heightened presence during the initial phases of blood feeding may be linked to the ticks' need to establish a secure attachment by means of the cement cone.

Secreted conserved proteins

Multiple proteomic and transcriptomic analyses of tick salivary glands have revealed the presence of numerous proteins that do not exhibit similarity to proteins found in other species within the public repositories. However, similar proteins have been identified in the saliva of various tick species and are referred to as secreted conserved proteins [11]. These proteins show the highest abundance in the initial saliva samples (groups Rm-1 to Rm-2) and gradually decrease towards the later samples (groups Rm-3 to FEF) (Fig. 3). These functionally unknown proteins comprise a significant fraction of the saliva of different tick species, suggesting a key biological role in tick biology. Further analyses using the TickSialoFam database have classified most of the proteins in these categories as follows: secreted conserved protein, secreted protein 28 kDa, secreted protein 13–14 kDa, secreted protein 23–24 kDa, 8.9 kDa family member, antigen 5/SCP domain-containing protein, tick salivary cytotoxin, ixodegrin, Mys-25–299, Mys-30–60, salivary protein one-of-each family, and salp15/ixostatin (Additional file 3:Table S3).

The 8.9-kDa family is exclusively found in hard ticks and exhibits a highly variable structure, with most forms containing one or two domains [59]. The only established function of the 8.9-kDa family is the inhibition of complement by binding to component C5 and preventing its activation, exemplified by the protein CirpT1 from Rhipicephalus pulchellus [86]. Using the crystal structure of CirpT1 complexed with C5 alone, a specific block of eight residues in the interaction interface was identified [59]. Among the five proteins classified within the 8.9-kDa family, Rm-3247 is the sole protein displaying this block of residues and appears to be the homolog of CirpT1 in R. microplus (Additional file 10: Fig. S7). Furthermore, the presence of proteins within the 8.9-kDa family exhibits a peak at the onset of the feeding period (Rm-1 to Rm-3), followed by a decline in group Rm-4, remaining relatively stable until the completion of the blood meal (Additional file 11: Fig. S8).

Proteins classified as cytotoxins are related to bacterial pore-forming proteins. The presence of these proteins in ticks suggests they were acquired by horizontal transfer [59], as was the case with the DAE antimicrobial proteins [87]. Nine of ten proteins annotated as cytotoxin are relatively more abundant at Rm-1 (Additional file 11: Fig. S8).

Ixostatin was the name given to a group of cysteine-rich protein sequences found in the sialotranscriptome of Ixodes pacificus. These sequences displayed remarkable similarities to the cysteine-rich domain of ADAMTS-4 (aggrecanase). The vast majority of ixostatins are found in the Ixodes genus, but a few sequences are also found in metastriate ticks. We have identified two sequences similar to Salp15 (Rm-19629 and Rm-66113), a protein isolated from I. scapularis, which inhibits CD4 + T cell activation [88].

The 13-kDa basic family is a small secretory family found in prostriate and metastriate ticks. This family is highly conserved among tick species and is characterized by six conserved cysteines. Alphafold2 predicts a compact globular structure formed by six alpha-helices. All structures present intact disulfide bonds, and the Dali search algorithm identified structural homologs among members of the odorant-binding, pheromone-binding, and venom 2 allergen families [59]. Odorant-binding proteins have not yet been found in ticks as prominent ligand scavengers involved in tick-host interactions, as this function is generally performed by lipocalins; however, a kratagonist function cannot be ruled out.

Immunity

The tick immune response is primarily based on innate immunity, with many compounds and mechanisms conserved across various species. While the hemolymph is the primary site of defense mechanisms in insects [89], the salivary glands express various immune response compounds that are secreted into the saliva [90, 91]. Consequently, in the saliva, numerous molecules with potential tick immunity-related activities have been detected. Specifically, proteins related to the immunity response are represented by microplusin and microplusin-like proteins, Toll-like receptors, alpha-2-macroglobulin-like proteins, evasins, ML domain-containing proteins, calreticulins, and DA P-36.

Microplusin is an antimicrobial peptide active against bacteria and fungi present in the tick R. microplus [92, 93] and, with a homologous peptide, in Amblyomma hebraeum [94]. The structure and biological function of microplusin have been characterized in the tick's hemolymph, ovaries, and eggs [92, 93]. The role of antimicrobial peptides in tick saliva may be associated with the prevention of microbial proliferation at the tick-feeding site.

The Toll-like pathway was originally discovered in Drosophila [95]. This highly conserved pathway presents in both invertebrates and vertebrates. It detects pathogen-associated molecular patterns (PAMPs), activating the immune response against various pathogens [96]. In ticks, the Toll-like system is not fully characterized, but the expression of its components is modulated by pathogen infection [89]. The precise nature of the presence of Toll-like compounds in saliva, whether they have a biological function or are a result of salivary gland degeneration, remains unclear. An interesting soluble form of Toll-like receptor displaying immunoregulatory functions has been found in human saliva [97].

Proteins with immunosuppressive activity clustered in the P36 family are immunosuppressive proteins identified in Dermacentor andersoni (DAP36) [98], H. longicornis (HL-p36) [99], Rhipicephalus haemaphysaloides (RH36) [100], and other ticks [101]. The suppressive activity was characterized in D. andersoni [102] and H. longicornis salivary glands [103]. Interestingly, in R. microplus, the removal of a Coxiella-like endosymbiont reduces the expression of the P36 protein, suggesting a complex host-microbiota-tick interaction [104].

Calreticulin is a Ca2+ binding protein with multifunctional properties. Initial studies showed it to be a molecular chaperone associated with the endoplasmic reticulum [105]. However, several studies have consistently shown that calreticulin has other roles as an intra- or extracellular protein [106], including in tick saliva [107,108,109]. Furthermore, studies have suggested a role for calreticulin as a potential biomarker of tick exposure in humans [110, 111] or an antigen for vaccination [112]. As a protein with immunomodulatory properties, it has been observed that calreticulin from various parasites, including Trypanosoma cruzi [113] and Hemonchus contortus [114], can inhibit the activation of the complement pathway and the blood clotting system. In contrast, the biological function of calreticulin in tick saliva remains unknown since these activities were not observed in A. americanum and R. microplus saliva [108, 109].

An insight into the core proteins related to different phases of feeding in Rhipicephalus microplus

To determine whether proteins were organized in clusters associated with adult ticks at different feeding stages, we subjected the normalized NSAF values to the CLICK algorithm within the expander program, which identified four protein clusters (Fig. 4).

Fig. 4
figure 4

Unsupervised clustering of proteins identified in the Rhipicephalus microplus saliva proteome throughout blood feeding. A The patterns were identified through clustering analysis using the CLICK algorithm. The x-axis represents the tick groups, the y-axis represents log2-based ratios standardized expression levels of cluster groups, and error bars represent the standard error. Each cluster is presented with triplicate samples. In parentheses, the number of proteins identified in each cluster is shown. B The frequency of the functional classes of proteins identified within each cluster is represented by the size of each sphere

Cluster 1 consisted of 130 proteins, primarily found in the Rm-1 group, representing the early stage of feeding. These proteins belong to various classes, including "Cytoskeletal," "Extracellular matrix/cell adhesion" (including glycine-rich proteins), "Heme/iron metabolism" (including the heme-binding protein HeLp), "Immunity" (including antimicrobial microplusin), "Secreted-lipocalin," "Metabolism" (carbohydrate, lipid, and energy), "Oxidant metabolism/detoxification," "Secreted-protease" (including mainly proteases from the M12B family), "Secreted-protease inhibitor" (including serpins RmS-7 and RmS-15), and "Secreted-secreted conserved protein" (including cytotoxins and a potential inhibitor of complement from the 8.9-kDa family) (Fig. 4b and Additional file 3: Table S3).

Cluster 2 included 29 proteins, which were predominantly present in groups Rm-2 and Rm-3. Proteins within this cluster include "Extracellular matrix/cell adhesion" (including mucin and glycine-rich protein), “Secreted-secreted conserved protein” (including 8.9-kDa family, ixodegrin, 23–24 kDa), "Immunity" (including microplusin-like and Toll-like), and "Secreted-lipocalin" (Fig. 4b and Additional file 3: Table S3).

Cluster 3 contained 18 proteins, mainly associated with the Rm-5 group. Proteins within this cluster include "Extracellular matrix/cell adhesion" (including mucin and glycine-rich protein), “Secreted-secreted conserved protein” (including 13 – 14 kDa, 23–24 kDa, and unknow), "Immunity" (including defensin and ML domain-containing protein), and "Secreted-lipocalin" (Fig. 4b and Additional file 3: Table S3).

Cluster 4 comprised 82 proteins, with their highest presence observed as ticks approached the end of feeding, as represented by groups Rm-8 and FEF (Fig. 4). These proteins belong to various classes, including "Metabolism" (carbohydrate, lipid, and intermediate), "Immunity" (including antimicrobial microplusin, evasin, ML domain-containing protein, and immunoglobulin G-binding protein), "Secreted-lipocalin," "Secreted-protease" (including mainly proteases from the A01, S01, and S10 family), "Secreted-protease inhibitor" (including cystatin, BmSEI-like, serpins RmS-3 and RmS-6, and TIL), and "Secreted-secreted conserved protein" (including 23–24 kDa, Salp15-like, and unknown) (Fig. 4b and Additional file 3: Table S3).

These results confirm the presence of a 'sialome switch' within R. microplus, facilitating the identification of proteins with similar expression patterns. Proteins presented within these clusters may represent the core of proteins needed at different phases of blood feeding.

An insight into the core proteins related to the sialome of different tick species

The identification of proteins in tick saliva and alterations in the saliva composition during the feeding process have been analyzed in various tick species. This includes proteome analysis of O. moubata males and females [30, 48], engorged females of R. sanguineus [49], partially and fully engorged female R. microplus [9], nymph and adult females of H. longicornis [8], and A. sculptum adult females [33]. Other studies have also performed analyses of saliva at different moments during tick feeding. In I. scapularis and A. americanum, saliva was analyzed at different day intervals [7, 31]. Additionally, the variation in saliva composition was analyzed in unfed I. scapularis and A. americanum ticks exposed to different hosts [32] and, in I. scapularis, nymphs infected and non-infected with B. burgdorferi [34].

To identify reciprocal best hits among tick saliva proteins described in the aforementioned studies, a BLASTp analysis was performed. We selected a final cutoff value of 1e−6 and required a coverage of at least 50%. In addition to the datasets of tick salivary proteins mentioned above, we included datasets of proteins identified in the cement of A. americanum [6] and I. scapularis [35].

Of the 284 proteins identified in this study, 267 have reciprocal best hits with salivary proteins identified in other tick species (Fig. 5 and Additional file 3: Table S3—sheet tab “Best-reciprocal”). These data suggest that ticks utilize a core set of functionally similar proteins that can regulate key host defense pathways to successfully feed. Functional classes conserved in the saliva of different tick species include hemelipoproteins, lipocalins, metalloproteases, protease inhibitors (serpins, cystatins, and TIL), and secreted conserved proteins, among others. While the functional roles of most of these salivary proteins remain to be determined, available evidence indicates that some of these proteins may regulate important tick feeding pathways, with potential variations among different tick species. For instance, the complement inhibitor CirpT1 is a member of the 8.9-kDa family and inhibits complement by binding to component C5, preventing its activation [86]. Best reciprocal hits showed these proteins are present in the saliva of most tick species. Interestingly, only homologs from R. microplus, A. americanum, and A. sculptum display the block of eight residues in the interaction interface of CirtT1 with the C5 protein (Additional file 12: Fig. S9), suggesting similar proteins in different tick species may act by inhibiting complement. Proteins identified in the saliva of I. scapularis lack these residues. In fact, I. scapularis has evolved with Salp20-like proteins that inhibit exclusively the alternative pathway of complement [115, 116]. This may represent that, although similar functional groups of salivary proteins are shared among different tick species, different strategies are used by metastriate and prostriate ticks to overcome host responses during blood feeding.

Fig. 5
figure 5

Venn diagram showing the number of shared orthologs in the Rhipicephalus microplus saliva proteome based on reciprocal best hits. The comparison is made with proteins identified in saliva of other tick species, including A Amblyomma americanum tick cement cone [6], B A. americanum saliva collected every 24 h for 8 days [7], C A. americanum saliva of tick exposed to different hosts [32], D A. sculptum saliva collected from unfed and partially fed ticks [33], E Haemaphysalis longicornis saliva collected from nymph and adult females [8], F Ixodes scapularis saliva collected from Borrelia burgdorferi-infected nymphs [34], G Ixodes scapularis tick cement cone [35], H I. scapularis tick saliva collected every 24 h for 5 days [31], I I. scapularis saliva of tick exposed to different hosts [32], J Ornithodoros moubata female and male saliva [30], and K R. microplus partially engorged female and fully engorged female saliva [9]

An insight into the host-derived proteins in the sialome of Rhipicephalus microplus

The presence of host-derived proteins or fragments of host-derived proteins was already described in several species as a true component of tick saliva and not as a contamination due to saliva harvesting procedures [9, 48, 49, 117]. Here, we describe the identification of 47 host-derived proteins as a component of the R. microplus saliva content. The NSAF levels for these host-derived proteins, which include serum albumin, fibrinogen, serotransferrin, cathelicidin, immunoglobulin G chains, and hemoglobin, among others are shown in Additional file 2: Table S2. These findings indicate that the occurrence of host-derived proteins and/or their fragments in tick saliva may represent a genuine and widespread recycling system employed by ticks. In our previous research, when comparing the saliva of partially and fully engorged ticks, we detected a significant number of heme-binding proteins derived from the host, such as serum albumin, hemopexin, apolipoprotein, and peroxiredoxin. In that study, we proposed a mechanism wherein, towards the end of feeding, the tick replaces hemelipoproteins with host-derived heme-carrier proteins [9]. However, in the current study, we only detected serum albumin as a host-derived heme-binding protein. Despite this finding, the presence of host-derived proteins in tick saliva strongly suggests a collaborative system between tick and host during blood feeding. The presence of fragments of host proteins could also be a mechanism to increase the saliva osmolality and consequently control the water content in the saliva. Further investigation is required to elucidate the role of these host-derived proteins in the tick-host relationship [118].

Conclusions

Hard ticks feed for prolonged periods, sometimes lasting several days or weeks, during which they must evade host responses to establish long-term parasitism. Proteome analysis has yielded significant insights into tick salivary proteins, revealing both the upregulation of various secreted proteins and the downregulation of others. This plasticity in salivary protein expression, often referred to as the 'sialome switch,' appears to be associated with the ticks' need to overcome various host responses to their feeding. Additionally, it may be linked to an immune evasion mechanism. The usual mechanisms that pathogens use to evade adaptive immunity are antigenic variation and immunomodulator proteins; both alter antigen recognition by host immune responses and allow the parasite to persist. Tick saliva collected at different times during the feeding process has many proteins related to immune evasion mechanisms. The findings presented in this study, along with the revealed sequences, enhance our comprehension of tick feeding biology and offer potential insights for discovering novel targets in the development of anti-tick strategies.

Availability of data and materials

The proteomic data have been deposited in the ProteomeXchange Consortium database via the PRIDE partner repository. To facilitate the exploration of this dataset, we developed a R shiny application that can be accessed online (https://tickproject.shinyapps.io/RmSaliva/). Additionally, a Windows-compatible hyperlinked Excel file that includes functional annotations, best reciprocal hits, and AlphaFold2 predictions for proteins identified in this study can be downloaded as a single.zip file from the following link: https://proj-bip-prod-publicread.s3.amazonaws.com/transcriptome/Rhip_microplus/Rm-saliva-2023/Table+S3.zip

References

  1. Kemp DH, Stone BF, Binnington KC. Tick Attachment and Feeding: Role of the Mouthparts, Feeding Apparatus, Salivary Gland Secretions and the Host Response. Physiology of Ticks [Internet]. Elsevier; 1982 p. 119–68. https://linkinghub.elsevier.com/retrieve/pii/B9780080249377500093. Accessed 21 Aug 2023.

  2. Neelakanta G, Sultana H. Tick saliva and salivary glands: what do we know so far on their role in arthropod blood feeding and pathogen transmission. Front Cell Infect Microbiol. 2022;11:816547.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Guo S, Dipietro LA. Factors affecting wound healing. J Dent Res. 2010;89:219–29.

  4. Francischetti IM, Sa-Nunes A, Mans BJ, Santos IM, Ribeiro JM. The role of saliva in tick feeding. Front Biosci. 2009;14:2051.

    Article  PubMed Central  CAS  Google Scholar 

  5. Nuttall PA. Tick saliva and its role in pathogen transmission. Wien Klin Wochenschr. 2023;135:165–76.

    Article  PubMed  Google Scholar 

  6. Hollmann T, Kim TK, Tirloni L, Radulović ŽM, Pinto AFM, Diedrich JK, et al. Identification and characterization of proteins in the Amblyomma americanum tick cement cone. Int J Parasitol. 2018;48:211–24.

    Article  PubMed  CAS  Google Scholar 

  7. Kim TK, Tirloni L, Pinto AFM, Diedrich JK, Moresco JJ, Yates JR, et al. Time-resolved proteomic profile of Amblyomma americanum tick saliva during feeding. Dinglasan RR, editor. PLoS Negl Trop Dis. 2020;14:e0007758.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Tirloni L, Islam MS, Kim TK, Diedrich JK, Yates JR, Pinto AFM, et al. Saliva from nymph and adult females of Haemaphysalis longicornis: a proteomic study. Parasites Vectors. 2015;8:338.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Tirloni L, Reck J, Terra RMS, Martins JR, Mulenga A, Sherman NE, et al. Proteomic analysis of cattle tick Rhipicephalus (Boophilus) microplus saliva: a comparison between partially and fully engorged females. PLoS ONE. 2014;9:e94831.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Chmelař J, Kotál J, Langhansová H, Kotsyfakis M. Protease inhibitors in tick saliva: the role of serpins and cystatins in tick-host-pathogen interaction. Front Cell Infect Microbiol. 2017;7:216.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Ribeiro JMC, Mans BJ. TickSialoFam (TSFam): a database that helps to classify tick salivary proteins, a review on tick salivary protein function and evolution, with considerations on the tick sialome switching phenomenon. Front Cell Infect Microbiol. 2020;10:374.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Martins LA, Bensaoud C, Kotál J, Chmelař J, Kotsyfakis M. Tick salivary gland transcriptomics and proteomics. Parasite Immunol. 2021. https://doi.org/10.1111/pim.12807.

  13. Ali A, Parizi LF, Ferreira BR, Vaz Junior I da S. Uma revisão sobre duas espécies distintas de Rhipicephalus: R. microplus e R. australis. Ciencia Rural. 2016;46.

  14. Friesen KJ, Kaufman WR. Salivary gland degeneration and vitellogenesis in the ixodid tick Amblyomma hebraeum: surpassing a critical weight is the prerequisite and detachment from the host is the trigger. J Insect Physiol. 2009;55:936–42.

    Article  PubMed  CAS  Google Scholar 

  15. Reck J, Berger M, Terra RMS, Marks FS, da Silva VI, Guimarães JA, et al. Systemic alterations of bovine hemostasis due to Rhipicephalus (Boophilus) microplus infestation. Res Vet Sci. 2009;86:56–62.

  16. Ciprandi A, de Oliveira SK, Masuda A, Horn F, Termignoni C. Boophilus microplus: its saliva contains microphilin, a small thrombin inhibitor. Exp Parasitol. 2006;114:40–6.

    Article  PubMed  CAS  Google Scholar 

  17. McDonald WH, Tabb DL, Sadygov RG, MacCoss MJ, Venable J, Graumann J, et al. MS1, MS2, and SQT—three unified, compact, and easily parsed file formats for the storage of shotgun proteomic spectra and identifications. Rapid Commun Mass Spectrom. 2004;18:2162–8.

    Article  PubMed  CAS  Google Scholar 

  18. Xu T, Park SK, Venable JD, Wohlschlegel JA, Diedrich JK, Cociorva D, et al. ProLuCID: an improved SEQUEST-like algorithm with enhanced sensitivity and specificity. J Proteomics. 2015;129:16–24.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  19. Tirloni L, Braz G, Nunes RD, Gandara ACP, Vieira LR, Assumpcao TC, et al. A physiologic overview of the organ-specific transcriptome of the cattle tick Rhipicephalus microplus. Sci Rep. 2020;10:18296.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Carvalho PC, Lima DB, Leprevost FV, Santos MDM, Fischer JSG, Aquino PF, et al. Integrated analysis of shotgun proteomic data with PatternLab for proteomics 4.0. Nat Protocols. 2016;11:102–17.

    Article  PubMed  CAS  Google Scholar 

  21. Karim S, Singh P, Ribeiro JMC. A deep insight into the sialotranscriptome of the gulf coast tick, Amblyomma maculatum. PLoS ONE. 2011;6:e28525.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Kim D-K, Kang B, Kim OY, Choi D, Lee J, Kim SR, et al. EVpedia: an integrated database of high-throughput data for systemic analyses of extracellular vesicles. J Extracell Vesicles. 2013;2:20384.

    Article  Google Scholar 

  23. Jia W, Chen S, Chi S, He Y, Ren L, Wang X. Recent progress on tick-borne animal diseases of veterinary and public health significance in China. Viruses. 2022;14:355.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Wickham H. ggplot2 [Internet]. Cham: Springer International Publishing; 2016. https://doi.org/10.1007/978-3-319-24277-4

  25. Hait TA, Maron-Katz A, Sagir D, Amar D, Ulitsky I, Linhart C, et al. The EXPANDER integrated platform for transcriptome analysis. J Mol Biol. 2019;431:2398–406.

    Article  PubMed  CAS  Google Scholar 

  26. Tang Y, Horikoshi M, Li W. ggfortify: unified interface to visualize statistical results of popular R packages. The R Journal. 2016;8:474.

    Article  Google Scholar 

  27. Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596:583–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Holm L, Kääriäinen S, Rosenström P, Schenkel A. Searching protein structure databases with DaliLite vol 3. Bioinformatics. 2008;24:2780–1.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  29. Moreno-Hagelsieb G, Latimer K. Choosing BLAST options for better detection of orthologs as reciprocal best hits. Bioinformatics. 2008;24:319–24.

    Article  PubMed  CAS  Google Scholar 

  30. Oleaga A, Carnero-Morán A, Valero ML, Pérez-Sánchez R. Proteomics informed by transcriptomics for a qualitative and quantitative analysis of the sialoproteome of adult Ornithodoros moubata ticks. Parasites Vectors. 2021;14:396.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Kim TK, Tirloni L, Pinto AFM, Moresco J, Yates JR, da Silva VI, et al. Ixodes scapularis tick saliva proteins sequentially secreted every 24 h during blood feeding. PLoS Negl Trop Dis. 2016;10:e0004323.

  32. Tirloni L, Kim TK, Pinto AFM, Yates JR, da Silva VI, Mulenga A. Tick-host range adaptation: changes in protein profiles in unfed adult Ixodes scapularis and Amblyomma americanum saliva stimulated to feed on different hosts. Front Cell Infect Microbiol. 2017;7:517.

  33. Esteves E, Maruyama SR, Kawahara R, Fujita A, Martins LA, Righi AA, et al. Analysis of the salivary gland transcriptome of unfed and partially fed Amblyomma sculptum ticks and descriptive proteome of the saliva. Front Cell Infect Microbiol. 2017;7:476.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Kim TK, Tirloni L, Bencosme-Cuevas E, Kim TH, Diedrich JK, Yates JR, et al. Borrelia burgdorferi infection modifies protein content in saliva of Ixodes scapularis nymphs. BMC Genomics. 2021;22:152.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Mulenga A, Radulovic Z, Porter L, Britten TH, Kim TK, Tirloni L, et al. Identification and characterization of proteins that form the inner core Ixodes scapularis tick attachment cement layer. Sci Rep. 2022;12:21300.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Hall TA. BioEdit: A user‐friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. 1999;41:95–8.

  37. Nicholas KB, Nicholas HBJr. GENEDOC: a tool for editing and annotating multiple sequence alignments. 1997;www.psc.edu/biomed/genedoc. www.psc.edu/biomed/genedoc

  38. Giachetto PF, Cunha RC, Nhani A, Garcia MV, Ferro JA, Andreotti R. Gene expression in the salivary gland of Rhipicephalus (Boophilus) microplus fed on tick-susceptible and tick-resistant hosts. Front Cell Infect Microbiol. 2020;9:477.

  39. Fernández-Ruiz N, Estrada-Peña A. Scenes from tick physiology: proteins of sialome talk about their biological processes. Front Cell Infect Microbiol. 2022;11:767845.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Maldonado-Ruiz LP, Montenegro-Cadena L, Blattner B, Menghwar S, Zurek L, Londono-Renteria B. Differential tick salivary protein profiles and human immune responses to lone star ticks (Amblyomma americanum) from the wild vs a laboratory colony. Front Immunol. 2019;10:1996.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Pham M, Underwood J, Oliva Chávez AS. Changing the recipe: pathogen directed changes in tick saliva components. IJERPH. 2021;18:1806.

  42. Karim S, Ribeiro JMC. An insight into the sialome of the lone star tick, Amblyomma americanum, with a glimpse on its time dependent gene expression. Schneider BS editor. PLoS ONE. 2015;10:e0131292.

    Article  PubMed  PubMed Central  Google Scholar 

  43. Mans BJ. Paradigms in tick evolution. Trends Parasitol. 2023;39:475–86.

    Article  PubMed  Google Scholar 

  44. Maya-Monteiro CM, Daffre S, Logullo C, Lara FA, Alves EW, Capurro ML, et al. HeLp, a heme lipoprotein from the hemolymph of the cattle tick, Boophilus microplus. J Biol Chem. 2000;275:36584–9.

    Article  PubMed  CAS  Google Scholar 

  45. Maya-Monteiro CM, Alves LR, Pinhal N, Abdalla DSP, Oliveira PL. HeLp, a heme-transporting lipoprotein with an antioxidant role. Insect Biochem Mol Biol. 2004;34:81–7.

    Article  PubMed  CAS  Google Scholar 

  46. Donohue KV, Khalil SMS, Mitchell RD, Sonenshine DE, Michael RR. Molecular characterization of the major hemelipoglycoprotein in ixodid ticks. Insect Mol Biol. 2008;17:197–208.

    Article  PubMed  CAS  Google Scholar 

  47. Gudderra NP, Sonenshine DE, Apperson CS, Roe RM. Tissue distribution and characterization of predominant hemolymph carrier proteins from Dermacentor variabilis and Ornithodoros parkeri. J Insect Physiol. 2002;48:161–70.

    Article  PubMed  CAS  Google Scholar 

  48. Díaz-Martín V, Manzano-Román R, Valero L, Oleaga A, Encinas-Grandes A, Pérez-Sánchez R. An insight into the proteome of the saliva of the argasid tick Ornithodoros moubata reveals important differences in saliva protein composition between the sexes. J Proteomics. 2013;80:216–35.

    Article  PubMed  Google Scholar 

  49. Oliveira CJ, Anatriello E, de Miranda-Santos IK, Francischetti IM, Sá-Nunes A, Ferreira BR, et al. Proteome of Rhipicephalus sanguineus tick saliva induced by the secretagogues pilocarpine and dopamine. Ticks and Tick-borne Diseases. 2013;4:469–77.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Valenzuela JG, Francischetti IMB, Pham VM, Garfield MK, Mather TN, Ribeiro JMC. Exploring the sialome of the tick Ixodes scapularis. J Exp Biol. 2002;205:2843–64.

    Article  PubMed  CAS  Google Scholar 

  51. Bozza MT, Jeney V. Pro-inflammatory actions of heme and other hemoglobin-derived DAMPs. Front Immunol. 2020;11:1323.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  52. Andersen JF, Ribeiro JMC. Salivary Kratagonists. Arthropod Vector: Controller of Disease Transmission, Volume 2 [Internet]. Elsevier; 2017, p. 51–63. https://linkinghub.elsevier.com/retrieve/pii/B9780128053607000046. Accessed 21 Aug 2023.

  53. Guizzo MG, Mans B, Pienaar R, Ribeiro JMC. A comparison of Illumina and PacBio methods to build tick salivary gland transcriptomes confirms large expression of lipocalins and other salivary protein families that are not represented in available tick genomes. Ticks Tick-borne Dis. 2023;14:102209.

    Article  PubMed  Google Scholar 

  54. Mans BJ. Chemical equilibrium at the tick-host feeding interface: a critical examination of biological relevance in hematophagous behavior. Front Physiol. 2019;10:530.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Francischetti IMB, Mather TN, Ribeiro JMC. Cloning of a salivary gland metalloprotease and characterization of gelatinase and fibrin(ogen)lytic activities in the saliva of the Lyme disease tick vector Ixodes scapularis. Biochem Biophys Res Commun. 2003;305:869–75.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  56. Ali A, Khan S, Ali I, Karim S, Da Silva VI, Termignoni C. Probing the functional role of tick metalloproteases: tick metalloproteases. Physiol Entomol. 2015;40:177–88.

    Article  CAS  Google Scholar 

  57. Decrem Y, Mariller M, Lahaye K, Blasioli V, Beaufays J, Zouaoui Boudjeltia K, et al. The impact of gene knock-down and vaccination against salivary metalloproteases on blood feeding and egg laying by Ixodes ricinus. Int J Parasitol. 2008;38:549–60.

    Article  PubMed  CAS  Google Scholar 

  58. Parizi LF, Githaka NW, Logullo C, Zhou J, Onuma M, Termignoni C, et al. Universal tick vaccines: candidates and remaining challenges. Animals. 2023;13:2031.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Mans BJ, Andersen JF, Ribeiro JMC. A deeper insight into the tick salivary protein families under the light of Alphafold2 and Dali: introducing the TickSialoFam 2.0 database. IJMS. 2022;23:15613.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Matsui T, Fujimura Y, Titani K. Snake venom proteases affecting hemostasis and thrombosis. Biochim Biophys Acta. 2000;1477:146–56.

    Article  PubMed  CAS  Google Scholar 

  61. Ekici ÖD, Paetzel M, Dalbey RE. Unconventional serine proteases: variations on the catalytic Ser/His/Asp triad configuration. Protein Sci. 2008;17:2023–37.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  62. Pichu S, Ribeiro JMC, Mather TN, Francischetti IMB. Purification of a serine protease and evidence for a protein C activator from the saliva of the tick, Ixodes scapularis. Toxicon. 2014;77:32–9.

    Article  PubMed  CAS  Google Scholar 

  63. Horn M, Nussbaumerová M, Šanda M, Kovářová Z, Srba J, Franta Z, et al. Hemoglobin digestion in blood-feeding ticks: mapping a multipeptidase pathway by functional proteomics. Chem Biol. 2009;16:1053–63.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  64. Franta Z, Frantová H, Konvičková J, Horn M, Sojka D, Mareš M, et al. Dynamics of digestive proteolytic system during blood feeding of the hard tick Ixodes ricinus. Parasites Vectors. 2010;3:119.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  65. Sorgine MHF, Logullo C, Zingali RB, Paiva-Silva GO, Juliano L, Oliveira PL. A heme-binding aspartic proteinase from the eggs of the hard Tick Boophilus microplus. J Biol Chem. 2000;275:28659–65.

    Article  PubMed  CAS  Google Scholar 

  66. Pohl PC, Sorgine MHF, Leal AT, Logullo C, Oliveira PL, Da Silva VI, et al. An extraovarian aspartic protease accumulated in tick oocytes with vitellin-degradation activity. Comp Biochem Physiol B: Biochem Mol Biol. 2008;151:392–9.

    Article  PubMed  Google Scholar 

  67. Huntington JA, Read RJ, Carrell RW. Structure of a serpin–protease complex shows inhibition by deformation. Nature. 2000;407:923–6.

    Article  PubMed  CAS  Google Scholar 

  68. Abbas MN, Chlastáková A, Jmel MA, Iliaki-Giannakoudaki E, Chmelař J, Kotsyfakis M. Serpins in tick physiology and tick-host interaction. Front Cell Infect Microbiol. 2022;12:892770.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  69. Gettins PGW. Serpin structure, mechanism, and function. Chem Rev. 2002;102:4751–803.

    Article  PubMed  CAS  Google Scholar 

  70. Coutinho ML, Bizzarro B, Tirloni L, Berger M, Freire Oliveira CJ, Sá-Nunes A, et al. Rhipicephalus microplus serpins interfere with host immune responses by specifically modulating mast cells and lymphocytes. Ticks Tick-borne Dis. 2020;11:101425.

    Article  PubMed  Google Scholar 

  71. Xu T, Lew-Tabor A, Rodriguez-Valle M. Effective inhibition of thrombin by Rhipicephalus microplus serpin-15 (RmS-15) obtained in the yeast Pichia pastoris. Ticks Tick-borne Dis. 2016;7:180–7.

  72. Tirloni L, Kim TK, Coutinho ML, Ali A, Seixas A, Termignoni C, et al. The putative role of Rhipicephalus microplus salivary serpins in the tick-host relationship. Insect Biochem Mol Biol. 2016;71:12–28.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  73. Cierpicki T, Bania J, Otlewski J. NMR solution structure of Apis mellifera chymotrypsin/cathepsin G inhibitor-1 (AMCI-1): structural similarity with Ascaris protease inhibitors. Protein Sci. 2000;9:976–84.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  74. Fogaça AC, Almeida IC, Eberlin MN, Tanaka AS, Bulet P, Daffre S. Ixodidin, a novel antimicrobial peptide from the hemocytes of the cattle tick Boophilus microplus with inhibitory activity against serine proteinases. Peptides. 2006;27:667–74.

  75. Sasaki SD, Tanaka AS. rBmTI-6, a Kunitz-BPTI domain protease inhibitor from the tick Boophilus microplus, its cloning, expression and biochemical characterization. Vet Parasitol. 2008;155:133–41.

    Article  PubMed  CAS  Google Scholar 

  76. Costa G, Silva F, Manzato V, Torquato RJ, Gonzalez Y, Parizi LF, et al. A multiepitope chimeric antigen from Rhipicephalus microplus-secreted salivary proteins elicits anti-tick protective antibodies in rabbit. Vet Parasitol. 2023;318:109932.

    Article  Google Scholar 

  77. Lu S, Da Rocha LA, Torquato RJS, Da Silva Vaz Junior I, Florin-Christensen M, Tanaka AS. A novel type 1 cystatin involved in the regulation of Rhipicephalus microplus midgut cysteine proteases. Ticks Tick-borne Dis. 2020;11:101374.

  78. Parizi LF, Sabadin GA, Alzugaray MF, Seixas A, Logullo C, Konnai S, et al. Rhipicephalus microplus and Ixodes ovatus cystatins in tick blood digestion and evasion of host immune response. Parasites Vectors. 2015;8:122.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Parizi LF, Githaka NW, Acevedo C, Benavides U, Seixas A, Logullo C, et al. Sequence characterization and immunogenicity of cystatins from the cattle tick Rhipicephalus (Boophilus) microplus. Ticks Tick-borne Dis. 2013;4:492–9.

  80. Lenarcic B, Turk V. Thyroglobulin type-1 domains in equistatin inhibit both papain-like cysteine proteinases and cathepsin D. J Biol Chem. 1999;274:563–6.

    Article  PubMed  CAS  Google Scholar 

  81. Zavašnik-Bergant T, Bergant MM. Exogenous thyropin from p41 invariant chain diminishes cysteine protease activity and affects il-12 secretion during maturation of human dendritic cells. Mattei F, editor. PLoS ONE. 2016;11:e0150815.

    Article  PubMed  PubMed Central  Google Scholar 

  82. Costa G, Silva F, Manzato V, Ricardo RJ, Gonzalez Y, Parizi LF, et al. A multiepitope chimeric antigen from Rhipicephalus microplus-secreted salivary proteins elicits anti-tick protective antibodies in rabbit. Vet Parasitol. 2023;318:109932.

    Article  Google Scholar 

  83. Hernández-Goenaga J, López-Abán J, Protasio AV, Vicente Santiago B, Del Olmo E, Vanegas M, et al. Peptides derived of Kunitz-type serine protease inhibitor as potential vaccine against experimental schistosomiasis. Front Immunol. 2019;10:2498.

  84. Tossi A, Sandri L, Giangaspero A. Amphipathic, α-helical antimicrobial peptides. Biopolymers. 2000;55:4–30.

    Article  PubMed  CAS  Google Scholar 

  85. Yue B. Biology of the extracellular matrix: an overview. J Glaucoma. 2014;23:S20–3.

    Article  PubMed  Google Scholar 

  86. Reichhardt MP, Johnson S, Tang T, Morgan T, Tebeka N, Popitsch N, et al. An inhibitor of complement C5 provides structural insights into activation. Proc Natl Acad Sci USA. 2020;117:362–70.

    Article  PubMed  CAS  Google Scholar 

  87. Chou S, Daugherty MD, Peterson SB, Biboy J, Yang Y, Jutras BL, et al. Transferred interbacterial antagonism genes augment eukaryotic innate immune function. Nature. 2015;518:98–101.

    Article  PubMed  CAS  Google Scholar 

  88. Anguita J, Ramamoorthi N, Hovius JWR, Das S, Thomas V, Persinski R, et al. Salp15, an Ixodes scapularis salivary protein, inhibits CD4+ T cell activation. Immunity. 2002;16:849–59.

    Article  PubMed  CAS  Google Scholar 

  89. Fogaça AC, Sousa G, Pavanelo DB, Esteves E, Martins LA, Urbanová V, et al. Tick immune system: what is known, the interconnections, the gaps, and the challenges. Front Immunol. 2021;12:628054.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Ali A, Zeb I, Alouffi A, Zahid H, Almutairi MM, Ayed Alshammari F, et al. Host immune responses to salivary components—a critical facet of tick-host interactions. Front Cell Infect Microbiol. 2022;12:809052.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  91. Wu J, Zhou X, Chen Q, Chen Z, Zhang J, Yang L, et al. Defensins as a promising class of tick antimicrobial peptides: a scoping review. Infect Dis Poverty. 2022;11:71.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  92. Esteves E, Fogaça AC, Maldonado R, Silva FD, Manso PPA, Pelajo-Machado M, et al. Antimicrobial activity in the tick Rhipicephalus (Boophilus) microplus eggs: cellular localization and temporal expression of microplusin during oogenesis and embryogenesis. Dev Comp Immunol. 2009;33:913–9.

  93. Fogaça AC, Lorenzini DM, Kaku LM, Esteves E, Bulet P, Daffre S. Cysteine-rich antimicrobial peptides of the cattle tick Boophilus microplus: Isolation, structural characterization and tissue expression profile. Dev Comp Immunol. 2004;28:191–200.

    Article  PubMed  Google Scholar 

  94. Lai R, Takeuchi H, Jonczy J, Rees HH, Turner PC. A thrombin inhibitor from the ixodid tick, Amblyomma hebraeum. Gene. 2004;342:243–9.

    Article  PubMed  CAS  Google Scholar 

  95. Anderson KV, Jürgens G, Nüsslein-Volhard C. Establishment of dorsal-ventral polarity in the Drosophila embryo: genetic studies on the role of the Toll gene product. Cell. 1985;42:779–89.

  96. Valanne S, Vesala L, Maasdorp MK, Salminen TS, Rämet M. The Drosophila toll pathway in innate immunity: from the core pathway toward effector functions. J Immunol. 2022;209:1817–25.

    Article  PubMed  CAS  Google Scholar 

  97. Zunt SL, Burton LV, Goldblatt LI, Dobbins EE, Srinivasan M. Soluble forms of Toll-like receptor 4 are present in human saliva and modulate tumour necrosis factor-α secretion by macrophage-like cells. Clin Exp Immunol. 2009;156:285–93.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  98. Bergman DK, Palmer MJ, Caimano MJ, Radolf JD, Wikel SK. Isolation and molecular cloning of a secreted immunosuppressant protein from Dermacentor andersoni salivary gland. J Parasitol. 2000;86:516–25.

    Article  PubMed  CAS  Google Scholar 

  99. Nakajima C, Silva Vaz ID, Imamura S, Konnai S, Ohashi K, Onuma M. Random sequencing of cDNA library derived from partially-fed adult female Haemaphysalis longicornis salivary gland. J Vet Med Sci. 2005;67:1127–31.

    Article  PubMed  CAS  Google Scholar 

  100. Wang X, Hou Y, Saha TT, Pei G, Raikhel AS, Zou Z. Hormone and receptor interplay in the regulation of mosquito lipid metabolism. Proc Natl Acad Sci USA. 2017;114:E2709–18.

    PubMed  PubMed Central  CAS  Google Scholar 

  101. Pichu S, Ribeiro JMC, Mather TN. Purification and characterization of a novel salivary antimicrobial peptide from the tick, Ixodes scapularis. Biochem Biophys Res Commun. 2009;390:511–5.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  102. Alarcon-Chaidez FJ, Muller-Doblies UU, Wikel S. Characterization of a recombinant immunomodulatory protein from the salivary glands of Dermacentor andersoni. Parasite Immunol. 2003;25:69–77.

    Article  PubMed  CAS  Google Scholar 

  103. Konnai S, Nakajima C, Imamura S, Yamada S, Nishikado H, Kodama M, et al. Suppression of cell proliferation and cytokine expression by HL-p36, a tick salivary gland-derived protein of Haemaphysalis longicornis. Immunology. 2009;126:209–19.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  104. Guizzo MG, Tirloni L, Gonzalez SA, Farber MD, Braz G, Parizi LF, et al. Coxiella endosymbiont of Rhipicephalus microplus modulates tick physiology with a major impact in blood feeding capacity. Front Microbiol. 2022;13:868575.

    Article  PubMed  PubMed Central  Google Scholar 

  105. Michalak M, Corbett EF, Mesaeli N, Nakamura K, Opas M. Calreticulin: one protein, one gene, many functions. Biochem J. 1999;344:281–92.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  106. Asgari S, Schmidt O. Is cell surface calreticulin involved in phagocytosis by insect hemocytes? J Insect Physiol. 2003;49:545–50.

    Article  PubMed  CAS  Google Scholar 

  107. Alarcon-Chaidez F, Ryan R, Wikel S, Dardick K, Lawler C, Foppa IM, et al. Confirmation of tick bite by detection of antibody to Ixodes calreticulin salivary protein. Clin Vaccine Immunol. 2006;13:1217–22.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  108. Alexandre Sanchez Ferreira C, Silva Vaz ID, Silva Da Silva S, Haag KL, Valenzuela JG, Masuda A. Cloning and partial characterization of a Boophilus microplus (Acari: Ixodidae) calreticulin. Exp Parasitol. 2002;101:25–34.

  109. Kim TK, Tirloni L, Radulovic Z, Lewis L, Bakshi M, Hill C, et al. Conserved Amblyomma americanum tick Serpin19, an inhibitor of blood clotting factors Xa and XIa, trypsin and plasmin, has anti-haemostatic functions. Int J Parasitol. 2015;45:613–27.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  110. Malouin R. Longitudinal evaluation of an educational intervention for preventing tick bites in an area with endemic lyme disease in Baltimore County, Maryland. Am J Epidemiol. 2003;157:1039–51.

    Article  PubMed  Google Scholar 

  111. Scott AL, Jaworski DC, Glass GE, Sanders ML, Schwartz BS, Needham GR, et al. Antibody to a cDNA-derived calreticulin protein from Amblyomma americanum as a biomarker of tick exposure in humans. Am J Trop Med Hyg. 1998;59:279–85.

    Article  PubMed  Google Scholar 

  112. Gao J, Luo J, Fan R, Fingerle V, Guan G, Liu Z, et al. Cloning and characterization of a cDNA clone encoding calreticulin from Haemaphysalis qinghaiensis (Acari: Ixodidae). Parasitol Res. 2008;102:737–46.

    Article  PubMed  Google Scholar 

  113. Ramírez G, Valck C, Molina MC, Ribeiro CH, López N, Sánchez G, et al. Trypanosoma cruzi calreticulin: A novel virulence factor that binds complement C1 on the parasite surface and promotes infectivity. Immunobiology. 2011;216:265–73.

    Article  PubMed  Google Scholar 

  114. Naresha S, Suryawanshi A, Agarwal M, Singh BP, Joshi P. Mapping the complement C1q binding site in Haemonchus contortus calreticulin. Mol Biochem Parasitol. 2009;166:42–6.

    Article  PubMed  CAS  Google Scholar 

  115. Hourcade DE, Akk AM, Mitchell LM, Zhou H, Hauhart R, Pham CTN. Anti-complement activity of the Ixodes scapularis salivary protein Salp20. Mol Immunol. 2016;69:62–9.

    Article  PubMed  CAS  Google Scholar 

  116. Valenzuela JG, Charlab R, Mather TN, Ribeiro JMC. Purification, cloning, and expression of a novel salivary anticomplement protein from the tick, Ixodes scapularis. J Biol Chem. 2000;275:18717–23.

    Article  PubMed  CAS  Google Scholar 

  117. Madden RD, Sauer JR, Dillwith JW. A proteomics approach to characterizing tick salivary secretions. Exp Appl Acarol. 2004;32:77–87.

    Article  PubMed  Google Scholar 

  118. Boulanger N, Wikel S. Induced transient immune tolerance in ticks and vertebrate host: a keystone of tick-borne diseases? Front Immunol. 2021;12:625993.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

This work utilized the computational resources of the NIH HPC Biowulf cluster (http://hpc.nih.gov). We acknowledge Anita Mora and Rose Perry-Gottschalk from the Visual and Medical Arts branch (RTB, NIAID, NIH) and Dr. Lucas C. Sousa-Paula (TPTU, NIAID, NIH) for figure preparation.

Funding

Open Access funding provided by the National Institutes of Health (NIH). LT and JMCR were supported by the Intramural Research Program of the National Institute of Allergy and Infectious Diseases, grant Tick saliva and its importance for tick feeding and pathogen transmission, Z01 AI001337-01 (LT), and Vector-Borne Diseases: Biology of Vector Host Relationship, Z01 AI000810-18 (JMCR). AM was supported by extramural Research Program of the National Institute of Allergy and Infectious Diseases (AI093858 and AI074789). This work was supported by CAPES grant #88882.346657/2019-01; CNPq #159522/2019-6, 405763/2018-2 and #465678/2014-9; INCT-EM (Instituto Nacional de Ciência e Tecnologia de Entomologia Molecular); FAPERJ #210.139/2023 and FAPERGS # 21/2551-0002221-3 (Brazil).

Author information

Authors and Affiliations

Authors

Contributions

ISVJ, AFMP, JKD, JRY, CT, and LT conceived and designed the study; ISVJ, SL, AFMP, JKD, JRY, CT, JMCR, and LT acquired and analyzed the data; ISVJ, SL, AFMP, JMCR, and LT interpreted the data; ISVJ and LT wrote the first draft; ISVJ, SL, AFMP, AM, CT, JMCR, and LT substantially revised subsequent drafts. All authors approved the final version of the manuscript.

Corresponding author

Correspondence to Lucas Tirloni.

Ethics declarations

Ethics approval and consent to participate.

This study was conducted according to the ethical and methodological aspects recommended by the International and National Directives and Norms of the Animal Experimentation Ethics Committee at UFRGS. The protocol (project no. 27559) was approved by the Comissão de Ética no Uso de Animais (CEUA) of UFRGS.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: Table S1.

Mass spectrometry data for tick proteins identified in the saliva of Rhipicephalus microplus. The summary (column B) outlines the validation criteria for protein identification within all groups of ticks. True proteins, identified in two out of the three technical replicates, are assigned a value of (1), while proteins detected in only one of the three runs were categorized as low confidence (0). This information is presented separately for groups Rm-1 (column C), Rm-2 (column S), Rm-3 (column AI), Rm-4 (column AY), Rm-5 (column BO), Rm-6 (column CE), Rm-7 (column CU), Rm-8 (column DK), and FEF (column EA). For each replicate, values representing unique peptides (UniquePeptideCount), peptide count (PeptideCount), spectral count (SpecCount), normalized spectral abundance factor (NSAF), and coverage percentage (Coverage) are provided.

Additional file 2: Table S2.

Mass spectrometry data for bovine proteins identified in the saliva of Rhipicephalus microplus. The summary (column B) outlines the validation criteria for protein identification within all groups of ticks. True proteins, identified in two of the three technical replicates, are assigned a value of (1), while proteins detected in only one of the three runs were categorized as low confidence (0). This information is presented separately for groups Rm-1 (column C), Rm-2 (column S), Rm-3 (column AI), Rm-4 (column AY), Rm-5 (column BO), Rm-6 (column CE), Rm-7 (column CU), Rm-8 (column DK), and FEF (column EA). For each replicate, values representing unique peptides (UniquePeptideCount), peptide count (PeptideCount), spectral count (SpecCount), normalized spectral abundance factor (NSAF), and coverage percentage (Coverage) are provided.

Additional file 3: Table S3.

A Windows-compatible hyperlinked Excel file that includes functional annotations, best reciprocal hits, and AlphaFold2 predictions for proteins identified in the proteome of Rhipicephalus microplus saliva. This file can be downloaded as a single.zip file from the following link: https://proj-bip-prod-publicread.s3.amazonaws.com/transcriptome/Rhip_microplus/Rm-saliva-2023/Table+S3.zip

Additional file 4: Fig. S1.

Expression patterns of the proteases identified in the Rhipicephalus microplus saliva proteome throughout blood feeding. Dots represent the average of NSAF values of each functional group with error bars denoting the standard error.

Additional file 5: Fig. S2.

Amino acid alignment (ClustalW) of proteins identified within the M12B, M13, and M17 families of metalloproteases identified in the Rhipicephalus microplus saliva proteome throughout blood feeding. The motif H-E-x(2)-H–x(2)-G-x(2)-H presented in metalloproteases is highlighted by a red box. The highly conserved residues are labeled in black, and the less conserved ones are in gray.

Additional file 6: Fig. S3.

Amino acid alignment (ClustalW) of proteins identified within the S01 family of serine-proteases identified in the Rhipicephalus microplus saliva proteome throughout blood feeding and the trypsinogen anionic precursor from Bos taurus (AA38513.1). The catalytic triad comprising Asp, His, and Ser is highlighted by asterisks. The highly conserved residues are labeled in black, and the less conserved ones are in gray.

Additional file 7: Fig. S4.

Amino acid alignment (ClustalW) of proteins identified within the C01 family of serine-proteases identified in the Rhipicephalus microplus saliva proteome throughout blood feeding and the cathepsin B from Homo sapiens (PDB 1GMY chain A). The catalytic dyad comprising Cys and His is highlighted by asterisks. The highly conserved residues are labeled in black, and the less conserved ones are in gray.

Additional file 8: Fig. S5.

Expression patterns of the protease inhibitors identified in the Rhipicephalus microplus saliva proteome throughout blood feeding. Each data point represents the average NSAF value for each protease family, with error bars denoting the standard error.

Additional file 9: Fig. S6.

Amino acid alignment (ClustalW) of proteins identified within the TIL family of protease inhibitors identified in the Rhipicephalus microplus saliva proteome throughout blood feeding and the chymotrypsin inhibitor from Apis mellifera (PDB 1CCV). The putative P1 residues are highlighted by asterisks. The highly conserved residues are labeled in black, and the less conserved ones are in gray.

Additional file 10: Fig. S7.

Amino acid alignment (ClustalW) of proteins identified within the 8.9 kDa family of secreted conserved proteins identified in the Rhipicephalus microplus saliva proteome throughout blood feeding and the tick complement inhibitor CirpT1 from Rhipicephalus pulchellus (PDB 6RPT). The specific block of eight residues in the interaction interface between CirpT1 and C5 was identified [59] and residues are highlighted by asterisks. The highly conserved residues are labeled in black, and the less conserved ones are in gray.

Additional file 11: Fig. S8.

Expression patterns of the secreted conserved proteins identified in the Rhipicephalus microplus saliva proteome throughout blood feeding. Each data point represents the average NSAF value for each protease family, with error bars denoting the standard error.

Additional file 12: Fig. S9.

Amino acid alignment (ClustalW) of CirpT1 homologs identified in the Rhipicephalus microplus saliva proteome and in the saliva proteome of other tick species and the tick complement inhibitor CirpT1 from Rhipicephalus pulchellus (PDB 6RPT). The specific block of eight residues in the interaction interface between CirpT1 and C5 was identified [59] and residues are highlighted by asterisks. The highly conserved residues are labeled in black, and the less conserved ones are in gray.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

da Silva Vaz Junior, I., Lu, S., Pinto, A.F.M. et al. Changes in saliva protein profile throughout Rhipicephalus microplus blood feeding. Parasites Vectors 17, 36 (2024). https://doi.org/10.1186/s13071-024-06136-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13071-024-06136-5

Keywords