SUMMARY
Acidocalcisomes are organelles conserved during evolution and closely related to the so-called volutin granules of bacteria and archaea, to the acidocalcisome-like vacuoles of yeasts, and to the lysosome-related organelles of animal species. All these organelles have in common their acidity and high content of polyphosphate and calcium. They are characterized by a variety of functions from storage of phosphorus and calcium to roles in Ca2+ signaling, osmoregulation, blood coagulation, and inflammation. They interact with other organelles through membrane contact sites or by fusion, and have several enzymes, pumps, transporters, and channels.
KEYWORDS: acidocalcisome, blood clotting, calcium, dense granules, polyphosphate
INTRODUCTION
Acidocalcisomes (ACs) are specialized organelles found in a variety of organisms and characterized by their acidic pH and high concentrations of calcium and polyphosphate (polyP). Initially identified in trypanosomatids, they have been related to the metachromatic or volutin granules, first reported in bacteria (1), and to the lysosome-related organelles (LROs) present in animal cells (2). These connections suggest that storage of polyP and calcium in acidocalcisomes or acidocalcisome-like organelles is a conserved property of both prokaryotes and eukaryotes.
A combination of microscopy techniques and biochemical analyses was initially used to identify and characterize these organelles in trypanosomatid parasites (3, 4). These were found to have a low pH sustained by a bafilomycin-sensitive proton pump (V-H+-ATPase), and high calcium, which was taken up by a P-type Ca2+-ATPase (4). X-ray microanalyses (5) showed that the organelles corresponded to the “inclusion vacuoles” (6) or “dense granules” (7, 8), previously identified to possess large quantities of phosphorus and cations in different trypanosomatids. Later work found that another proton pump, a vacuolar proton pyrophosphatase, V-H+-PPase, was also present in these trypanosomatid organelles (9) and could be used as a suitable marker for their isolation (10). 31P nuclear magnetic resonance (NMR) (11, 12) and biochemical analyses (13) uncovered the presence of large amounts of polyP. In the meantime, investigation of other cells known to possess dense granules or a V-H+-PPase led to the discovery of similar organelles in other eukaryotic supergroups besides those belonging to the Discoba (trypanosomatids) (14–16), like Amoebozoa (Dictyostellium discoideum) (17), Archaeplastida (Chlamydomonas reinardtii) (18), Stramenopila-Alveolata-Rizaria [like Toxoplasma gondii (19), Plasmodium spp. (20), and Eimeria spp. (21)], and Opisthokonta {including sea urchins (22), insects (23), and humans [platelets (24) and mast cells (25)]}.
Finally, proteomic analyses of trypanosomatid acidocalcisomes allowed for the identification of numerous transporters, pumps and channels, and some enzymes that made these organelles so peculiar, and defined the mechanisms they use for Ca2+ signaling, synthesis, and degradation of polyP, phosphate (Pi) release, and accumulation of organic and inorganic cations (26). As expected, it was found that acidocalcisomes and acidocalcisome-like organelles of different species have distinct biochemical compositions, but all have in common the storage of polyP and calcium. The organelle is probably the earliest acidic calcium store that appeared during evolution (27). The study of acidocalcisomes by different groups has shed light on the important roles that these organelles play in cellular processes (28), and these will be the subject of this review.
PHYLOGENETIC DISTRIBUTION OF ACIDOCALCISOMES
The calcium and polyP storage capacity of acidocalcisomes is also present in the prokaryotic subcellular compartments known as polyP granules. These granules were reported by the end of the 19th century and named metachromatic (1) or volutin (29) granules. These names represented their property of staining red or purple when treated with toluidine blue (metachromatic granules) and the fact that they were first detected in Spirillum volutans (volutin granules). Staining with toluidine blue was applied to numerous protists that were then reported to have volutin granules, like yeasts (30, 31) and trypanosomatids (32). The finding that the number of granules within the yeast vacuoles increased when polyP increased led to a change in their name to polyP granules (30). Both bacteria (1, 29) and archaea (33, 34) possess polyP granules.
Early reports suggested that bacterial polyP granules were surrounded by a limiting membrane (35), but a membrane was not detected in other bacteria with current microscopy techniques, and it has been suggested that they resemble membrane-less compartments arising by liquid-liquid phase separation (36, 37). In this regard, polyP has been shown to induce liquid phase separation of proteins (38). However, some bacteria, like Agrobacterium tumefaciens (39) and Rhodospirillum rubrum (40), have internal membrane-bound vesicles as detected by electron microscopy (EM) of intact bacteria and subcellular fractions. Vesicles were also found in A. tumefaciens after they were snap frozen in living cells and then fractured and etched (QFDEEM) (41). Some of these vesicles, which are also detected by staining with dyes that accumulate in acidic compartments (cycloprodigiosin, lysosensor DND 167), are labeled with antibodies against the V-H+-PPase, as determined by immunofluorescence and electron microscopy analyses (39, 40). Subcellular fractionation found co-localization of the V-H+-PPase activity with polyP as determined by biochemical methods, suggesting that they were acidocalcisome-like organelles (39, 40).
More recent work in A. tumefaciens during different phases of growth proposed that acidocalcisomes and polyP granules are different subcellular structures (42). However, some contradictions were reported. Globular dark structures, detected by bright field microscopy, were shown to co-localize with either 4′,6-diamidino-2-phenylindole (DAPI)-fluorescent foci (polyP granules) [Fig. 4 of reference (42)] or the cell pole-located V-H+-PPase-GFP fluorescence [Fig. 5D and E of reference (42)], or with MitoTracker [Fig. 5B of reference (42)], used to detect acidocalcisomes, supporting that they represent the same structures. When the co-localization of DAPI and tagged V-H+-PPase was investigated, the authors found higher expression of the V-H+-PPase tagged with GFP in vesicles during the late stationary phase when DAPI-stainable polyP was rarely present. In the rare cases when both were detected, they were located closely together (42). DAPI cannot stain very short polyP (43, 44), and some acidocalcisomes are known to be rich in this type of polyP [tripolyphosphate (polyP3), polyP4, and polyP5] (11), which could explain the lack of detection of DAPI-positive vesicles in late stationary phase. It would be interesting to investigate whether this could be the reason for the apparent lack of detection of co-localization of both the V-H+-PPase and polyP in some cases.
Interestingly, sulfide-oxidizing bacteria of the genus Beggiatoa accumulate polyP and calcium in acidocalcisome-like inclusions surrounded by a lipid layer, although they are not notably acidic (45).
Acidocalcisomes with characteristics similar to those found initially in Trypanosoma brucei (4) and Trypanosoma cruzi (3) were identified in all trypanosomatids investigated, including digenetic Leishmania mexicana, Leishmania donovani (46, 47), Trypanosoma evansi (48), and Phytomonas francai (16), as well as in monogenetic trypanosomatids (49) and Naegleria gruberi (41), which also belong to the Discoba “supergroup” of eukaryotes.
Most other major supergroups of eukaryotes have species containing acidocalcisomes. One of the first organisms investigated was Toxoplasma gondii (19), which belongs to the Stramenopila-Alveolata-Rizaria supergroup. These acidic Ca2+ stores were biochemically characterized (19). A V-H+-PPase was identified (50) and shown to co-localize with a Ca2+-ATPase (51). The V-H+-PPase was needed for polyP storage (52). The organelles were isolated (53) and were shown to transport Ca2+ and protons (54). Other Alveolata members such Plasmodium falciparum, (55), Plasmodium berghei (20), and Eimeria spp. (21) also possess acidocalcisomes. The presence of polyP in the acidocalcisome-like organelles of ciliates has not been investigated (56, 57).
Within Archaeplastida, which includes algae and plants, the green alga Chlamydomonas reinhardtii was known to possess polyP granules within “acidic vacuoles” (58) and a protein that reacted with antibodies against a plant V-H+-PPase and localized to contractile vacuoles, the Golgi complex, and “intermediate-sized vesicles” (59). The measurement of V-H+-PPase activity, which co-localized with a V-H+-ATPase in vacuoles containing polyP determined that these acidic vacuoles or intermediate-sized vesicles were acidocalcisomes (18). Interestingly, although the vacuoles of higher plants also possess a V-H+-PPase and a V-H+-ATPase and several transporters with orthologs present in acidocalcisomes of several species, efforts to find polyP in higher plants have been unsuccessful (60). Other species of Chlamydomonas (41) and the green algae Dunaliella salina (61) and Desmodesmus sp. (62) possess polyP in acidic vacuoles compatible with acidocalcisomes. The red alga Cyanidioschyzon merolae was found to possess polyP in vacuoles with morphology similar to that of acidocalcisomes (63). These were isolated using iodixanol gradient centrifugation and shown to contain several pumps (V-H+-ATPase and V-H+-PPase) and transporters (for Zn2+ and for Fe2+/Mn2+) also present in trypanosomatid acidocalcisomes (63).
The supergroup Amoebozoa includes Dictyostelium discoideum, which was known to possess “mass-dense granules” rich in phosphorus and calcium (64, 65). Mass-dense granules have polyP, a V-H+-ATPase, and a Ca2+-ATPase (Pat1), as occurs with other acidocalcisomes, and a pyrophosphatase with similarities to the V-H+-PPase (17). In addition, a membrane-bound pyrophosphatase had been described before in D. discoideum (66). A protein was detected with heterologous antibodies against the plant V-H+-PPase. However, the pyrophosphatase activity was not completely inhibited by the inhibitors aminomethylenediphosphonate (AMDP) and imidodiphosphate (IDP), and no PPi-driven proton transport was reported (17), suggesting that this was not a typical V-H+-PPase, in agreement with genetic evidence of the absence of the gene in the genome of D. discoideum (41). Another amoebozoan, Entamoeba histolytica, also possesses electron-dense granules rich in Pi, PPi, and cations (Na+, Mg2+, K+, Ca2+, and Fe2+) (67), but the presence of polyP in these acidocalcisome-like organelles has not been investigated.
The supergroup Opisthokonta includes fungi and animals and has species possessing acidocalcisome-like organelles. The yeast vacuole has been proposed as an acidocalcisome-like organelle (68) as it is acidic (it has a V-H+-ATPase but not a V-H+-PPase), rich in polyP and cations, and has several transporters in common with those of trypanosomatid acidocalcisomes. Other fungi, such as Candida spp. (69, 70), Neurospora crassa (71), and arbuscular mycorrhizal fungi (72, 73), have a vacuolar compartment in the form of small vesicles or tubules rich in polyP and with similarities to acidocalcisomes. Acidocalcisome-like vacuoles were also found in the yolk of insect (23) and chicken (74) eggs and in sea urchin eggs (22). In all these cases, these vacuoles are acidic and rich in polyP and calcium. In humans, it was found that some lysosome-related organelles such as platelet-dense granules (24) and mast cell granules (25) are acidic, rich in Pi, PPi, polyP, and calcium and similar to acidocalcisomes.
The evidence reported indicates that acidocalcisomes and acidocalcisome-like vacuoles are widely distributed and reveals the importance of polyP during evolution.
STRUCTURAL CHARACTERISTICS OF ACIDOCALCISOMES
Acidocalcisomes were first identified in trypanosomes by their staining with acridine orange that was prevented by the V-H+-ATPase inhibitor bafilomycin A1 or the K+/H+ exchanger nigericin (3, 4). Later work (5, 14) demonstrated that the acidic vacuoles of Trypanosoma cruzi and T. brucei corresponded to the electron-dense vacuoles detected by electron microscopy (Fig. 1A through C) that, when analyzed by X-ray microanalyses, were found to be rich in cations (Ca2+, Mg2+, K+, Zn2+, and Fe2+), oxygen, and phosphorus (7) (Fig. 1D). Similar electron-dense compartments had been described before in Trypanosoma cyclops (8) and Leishmania major (6).
Fig 1.
Ultrastructure of acidocalcisomes. (A and B) Thin sections of Leishmania amazonensis (A) and Herpetomonas anglusteri (B) submitted to conventional transmission electron microscopy. Acidocalcisomes are partially filled with electron-dense material (“granule”) (arrows) or appear empty or with an electron-dense periphery (arrowheads). Bars are 280 nm (A) or 130 nm (B). (C) transmission electron microscopy of whole bloodstream trypomastigotes of T. brucei. The arrowheads indicate acidocalcisomes. Bar, 2 µm. (D) X-ray microanalysis spectrum of dense organelles in whole bloodstream trypomastigotes. (Panels A and B were reproduced from reference 49 with permission from Elsevier; panels C and D were reproduced from reference 14 with permission from Taylor & Francis Informa UK Ltd.) ER, endoplasmic reticulum; G, glycosome; M, mitochondrion.
The use of electron microscopy led to the identification of the acidocalcisomes by conventional electron microscopy as “empty” vacuoles sometimes surrounded by a peripheral dense region or showing interior electron-dense granules (75) (Fig. 1A and B). The appearance of empty vacuoles is due to the procedure applied for electron microscopy preparation that depletes their content. By cryo-electron microscopy, however, acidocalcisomes appear as electron-dense spheres (5). Depositing these protists on a grid, letting them dry, and observing them by EM also allow their detection in intact form as electron dense spheres (5) (Fig. 1C). Their size has been measured in different trypanosomatids (16, 49, 75) and varies from ~50 to ~250 nm in diameter reaching values of ~600 nm in diameter in Leishmania spp. (76). They occupy ~2% of the total cell volume (49). Some acidocalcisomes are elongated, like in Phytomonas spp. (16, 49), or pleomorphic, such as those in Leishmania mexicana amazonensis, grown in a deficient culture medium (76). There is no apparent specific distribution in the cells and their numbers are variable. In T. brucei bloodstream forms, there are about 40 acidocalcisomes per cell at the start of the cell cycle and their numbers increase to ~56 prior to cell division (77).
The ultrastructural analysis of acidocalcisomes of the alga Chlamydomonas reinhardtii and other protists was undertaken using living cells snap-frozen at liquid helium temperatures, subjected to freeze-fracture, deep etching, and platinum rotary-replication, and observed by transmission electron microscopy (41). This technique allowed the identification of a large population of intramembranous particles in the P-fracture concave face of the membrane (contiguous to the cytoplasm) and attributed them to intramembrane proteins (Fig. 2). The convex E-fracture face (contiguous to the lumen) was smooth or rugose, depending on its exposure to etching. An interior granule, attributed to polyP in a gel configuration, was detected in stationary phase cells or in those submitted to nitrogen (N)-starvation (41) (Fig. 2). The authors also investigated the distribution of V-H+-PPase-encoding genes in eukaryotes, finding three clades and that clade-1 proteins (K+-stimulated) are acidocalcisome associated (41).
Fig 2.

Cross-fractured acidocalcisomes. Two acidocalcisomes contiguous to one another from log phase C. reinhardtii N-starved for 24 h. The fracture plane exposed the P-face of the right organelle and the E-face of the left organelle. A large population of intramembranous particles is in the P-face of the membrane. The convex E-fracture face is rugose. “Rim” denotes the outer edge. Bar, 100 nm. P, P-fracture; E, E-fracture; g, polyphosphate granule. (Reproduced from reference 41 with permission from Elsevier.)
Acidocalcisomes have also been analyzed by cryofixation of cells and scanning transmission electron microscopy tomography combined with elemental mapping using a high-performance setup of X-ray detectors (78). The work reported that the elemental distribution was not homogenous but rather organized in nanodomains within the organelles with cationic elements displaying a self-excluding pattern (78).
In summary, the structural characteristics of acidocalcisomes are quite peculiar and different from any other organelle.
CHEMICAL COMPOSITION OF ACIDOCALCISOMES
The acidocalcisome matrix is rich in Pi, PPi, and polyP complexed with organic and inorganic cations and has a few enzymes. The phosphorus compounds have been identified by 31P NMR of isolated organelles from different trypanosomatids (11), and apicomplexan (79) parasites, or by labeling some of these cells with 32Pi (13). Very short-chain polyP (polyP3, PolyP4, and polyP5) was detected by 31P NMR in trypanosomatids (11). The results suggested that the average chain length of polyP is of ~3.2 phosphates. Longer-chain polyP was more difficult to identify. However, using polyacrylamide gel electrophoresis analyses of isolated acidocalcisomes from T. brucei it was possible to also identify the presence of short-chain (<300 mer) and long-chain (>300–1,000 mer) polyP (80).
The presence of organic cations, mainly basic amino acids like arginine, lysine, and ornithine, was demonstrated in acidocalcisomes from T. cruzi epimastigotes, where they represent about 90% of the total amino acid pool (81). Although the presence of polyamines has not been reported in acidocalcisomes of trypanosomatids, they possess polyamine transporters (26) as the acidocalcisome-like vacuoles of Saccharomyces cerevisiae, which are rich in these cationic compounds (82, 83).
Inorganic cations (Ca2+, Mg2+, Na+, K+, and Zn2+) are also abundant in acidocalcisomes of most species investigated by X-ray microanalysis methods. Chlamydomonas reinhardtii acidocalcisomes are also rich in copper (84), manganese (85), and iron (86), while iron has been found in acidocalcisomes of Phytomonas spp., T. cruzi, and other trypanosomatids (16, 49, 75, 76). Some metals, like manganese, iron, and copper, could be present in smaller amounts in some organisms, below the detection limit of X-ray microanalysis, as these cells possess transporters for them. Liquid chromatography (LC) studies of the acidocalcisome-like vacuoles of yeast have demonstrated that most of the iron, zinc, and manganese (and some copper) are coordinated to polyP chains (87). The results indicated that these metals are present in vacuoles in low-molecular-mass forms that migrate according to masses ranging from 500 to 10,000 Da. The observed LC peaks comigrated with phosphorus, supporting the presence of metal-polyP complexes. Treatment with polyphosphatase disrupted these complexes. In some cases, the metals comigrated with each other, suggesting that multiple metals either bind to the same chain or to different chains of about the same length. The length distribution of polyP chains in the samples was of chains of 6–20 units long, suggesting that these lengths were either especially stable or produced at especially high rates. The authors suggested that these preferences must reflect the steady-state distribution resulting from opposing polyP synthesis and degradation processes or that metals may also preferentially bind to and/or stabilize chains of specific lengths (87).
Isolated acidocalcisomes from different trypanosomatids (T. brucei, T. cruzi, and Leishmania major) were also studied by magic-angle spinning 31P NMR spectroscopy and resulted in the detection of condensed phosphates with the dominant presence of polyP3 and low abundance of long-chain polyP (12).
Lipid analysis of highly purified acidocalcisomes from T. cruzi (88) showed very low amounts of 3β-hydroxysterols. Alkylacyl phosphatidylinositol (16:0/18:2), diacyl phosphatidylinositol (18:0/18:2), diacyl phosphatidylcholine (16:0/18:2; 16:1/18:2; 16:2/18:2; 18:1/18:2, and 18:2/18:2), and diacyl phosphatidylethanolamine (16:0/18:2 and 16:1/18:2) were identified by electrospray ionization-mass spectrometry (88). A glycoinositolphospholipid (GIPL) was also detected with a structure apparently different from the GIPL found in microsomal fractions (88).
Proteomic analysis of T. brucei acidocalcisomes (26) and N-terminal or C-terminal tagging of proteins by the TrypTag.org project (89, 90) were able to identify more than 30 proteins. The acidocalcisome localization was validated for some of them by immunofluorescence co-localization with the V-H+-PPase (Table 1; Fig. 3). Figure 4 shows the scheme of a T. brucei acidocalcisome.
TABLE 1.
Acidocalcisome localization of T. brucei proteins, validated by marker co-localization
| Gene | Name | TMDd | MWe | References |
|---|---|---|---|---|
| Tb927.11.11160 | Phosphate transporter 91 (fragment) (PHO91) | 8 | 81.451 | Huang et al. (26) Billington et al. (89)a |
| Tb927.9.10340 | Polyamine transporter 1 | 11 | 54.065 | Huang et al. (26) |
| Tb927.3.800 | Vacuolar iron transporter | 3 | 30.573 | Huang et al. (26) Billington et al. (89)a |
| Tb927.4.4960 | Zn2+ transporter (ZnT1) |
5 | 50.697 | Huang et al. (26) |
| Tb927.8.7460 | Zn2+ transporter (ZnT2) |
5 | 50.823 | Billington et al. (89)a |
| Tb927.9.5490 | Cation transporter (Mg2+) | 1 | 75.937 | Billington et al. (89)a |
| Tb927.5.1260 | Sulfate transporter | 11 | 62.563 | Billington et al. (89)a |
| Tb927.11.16630 | Major facilitator superfamily MFS,b nodulin-like | 14 | 74.876 | Billington et al. (89)a |
| Tb927.11.12270 | MFS, drug resistance protein | 10 | 53.057 | Billington et al. (89)a |
| Tb927.10.12400 | Multidrug and Toxic extrusion protein | 11 | 51.410 | Billington et al. (89)a |
| Tb927.7.3900 | Vacuolar transporter chaperone 1 | 3 | 19.768 | Fang et al. (91) |
| Tb927.11.12220 | Vacuolar transporter chaperone 4 | 2 | 91.362 | Lander et al. (92) Ulrich et al. (93) Huang et al. (26) Billington et al. (89)a |
| Tb927.11.1260 | Cu-ATPase (ATP7) | 8 | 102.336 | Isah et al. (94); Paul et al. (95) Billington et al. (89)a |
| Tb927.8.1160 | Ca2+-ATPase (PMC1) | 10 | 121.242 | Luo et al. (96) |
| Tb927.4.4380 | Vacuolar H+-PPase 1 | 16 (SP) | 85.936 | Rodrigues et al. (14) Billington et al. (89)a |
| Tb927.8.7980 | Vacuolar H+-PPase 2 | 16 | 86.004 | Rodrigues et al. (14) Billington et al. (89)a |
| Tb927.5.1300 | Vacuolar H+-ATPase (subunit a) | 6 | 89.631 | Vercesi et al. (4) Huang et al. (26) |
| Tb927.5.550 | Vacuolar H+-ATPase (subunit D) | – | 42.849 | Vercesi et al.(4) Huang et al. (26) |
| Tb927.8.2310 | Vacuolar H+-ATPase (subunit G) | – | 12.749 | Vercesi et al. (4) Billington et al. (89)a |
| Tb927.4.1080 | Vacuolar H+-ATPase (subunit A) | – | 67.750 | Vercesi et al. (4) Billington et al. (89)a |
| Tb927.11.11690 | Vacuolar H+-ATPase (subunit B) | – | 55.620 | Vercesi et al. (4) Billington et al. (89)a |
| Tb927.8.2770 | Inositol trisphosphate receptor | 5 | 342.483 | Huang et al. (97) Billington et al. (89)a |
| Tb927.11.12490 | Potassium channel (IRK) | 3 | 61.904 | Steinmann et al. (98) Billington et al. (89)a |
| Tb927.11.7080 | Vacuolar soluble pyrophosphatase | – | 47.297 | Lemercier et al. (99) Huang et al. (26) |
| Tb927.10.7020 | Acid phosphatase | – | 49.918 | Huang et al. (26) |
| Tb927.6.4630 | Kinetoplastid-specific phosphoprotein phosphatase (Ppn2) | 1 | 39.849 | Huang et al. (26) Billington et al. (89)a |
| Tb927.11.10650 | AP-3c β3 subunit | – | 100.360 | Huang et al. (100) |
| Tb927.5.3610 | AP-3 δ subunit | 1 | 125.404 | Huang et al. (100) |
| Tb927.10.10800 | Palmitoyl acyl transferase 2 | 3 | 49.905 | Huang et al. (26); Billington et al. (89)a Emmer et al. (101) |
| Tb927.10.6180 | FLA1-like protein | 1 | 54.878 | Huang et al. (26) Sun et al. (102) Billington et al. (89)a |
| Tb927.10.9560 | Oxidoreductase | – | 36.100 | Billington et al. (89)a |
| Tb927.11.1890 | Serine threonine kinase-associated teceptor | – | 36.025 | Billington et al. (89)a |
| Tb927.4.860 | Hypothetical protein | – | 33.145 | Billington et al. (89)a |
| Tb927.4.5335 | Hypothetical protein | – | 10.705 | Billington et al. (89)a |
| Tb927.8.6990 | Hypothetical protein | – | 10.759 | Billington et al. (89)a |
| Tb927.9.12070 | Hypothetical protein | – | 109.870 | Billington et al. (89)a |
| Tb927.10.11810 | Hypothetical protein | – | 69.988 | Billington et al. (89)a |
| Tb927.11.8250 | Hypothetical protein | 1 | 19.345 | Billington et al. (89)a |
Tentative localization based on immunofluorescence analysis of C-terminal or N-terminal tagged proteins (TrypTag project).
MFS, major facilitator superfamily.
AP-3, adaptor protein 3.
TMD, transmembrane domain.
MW, molecular weight.
Fig 3.
Immunofluorescence analysis of T. brucei. (A) TbIP3R co-localizes with the V-H+-PPase in acidocalcisomes of procyclic (PCF) trypomastigotes (Pearson’s correlation coefficient of 0.8399). Yellow in merge images indicates co-localization. Scale bars: 10 µm. (B) Western blot analysis of TbIP3R expressed in PCF trypanosomes using polyclonal anti-TbIP3R antibody. Lysate containing 30 mg of protein from PCF trypanosomes was subjected to SDS/PAGE on 4%–15% polyacrylamide gel and transferred to a nitrocellulose membrane. Molecular weight markers on the left and arrow shows the band corresponding to TbIP3R. (Reproduced from reference 26.) TbIP3R, T. brucei IP3R.
Fig 4.
Scheme of an acidocalcisome of T. brucei with validated transporters and channels. Uptake of Ca2+ is through a Ca2+-ATPase, and Ca2+ release is by the IP3 receptor. H+ is pumped in by either the V-H+-PPase or the V-H+-ATPase. A vacuolar iron transporter (VIT1) can transport either Mn2+ or Fe2+, and two Zn2+ transporters (ZnT1 and ZnT2, not shown) transport Zn2+. There is a PA transporter and K+ channel. A VTC complex synthesizes polyP using ATP and translocates it into the organelle. A Na+/Pi symporter (Pho91) releases Na+ and Pi. Within acidocalcisomes, there is a VSP with pyrophosphatase and exopolyphosphatase activity. Several adaptor protein 3 complex subunits localize to the acidocalcisome (not shown). PA, polyamine; VSP, vacuolar soluble pyrophosphatase; VTC, vacuolar transporter chaperone.
T. cruzi has orthologs to most acidocalcisome proteins of T. brucei, and their localization was in part determined by proteomic analysis (103). Co-localization studies validated some of them. A few proteins with no acidocalcisome orthologs in T. brucei have also been found in acidocalcisomes of T. cruzi (Table 2).
TABLE 2.
Validated acidocalcisome proteins present in T. Cruzia
| Gene | Name | TMDb | MWc | References |
|---|---|---|---|---|
| TcCLB.511439.50 | Zn2+ transporter 1 | 5 | 49.901 | Ferella et al. (103) |
| TcCLB.511127.100 | Vacuolar transporter chaperone 4 | 2 | 91.558 | Ulrich et al. (93) |
| TcCLB.506401.170 | Ca2+-ATPase (PMC1) | 8 | 121.912 | Lu et al. (104) |
| TcCLB.510773.20 | Vacuolar H+-PPase 1 | 16 | 85.338 | Scott et al. (9) |
| TcCLB.508257.40 | Aquaporin 1 | 6 | 24.711 | Montalvetti et al. (105) Rohloff et al. (106) |
| TcCLB.509461.90 | Inositol trisphosphate receptor | 5 | 337.362 | Lander et al. (107) Chiurillo et al. (108) |
| TcCLB.503613.60 | Vacuolar soluble pyrophosphatase | – | 47.854 | Gallizi et al. (109) |
| TcCLB.506311.20 | Palmitoyl acyl transferase 2 | 4 | 51.182 | Batista et al. (110) |
| TcCLB.506247.220 | Histidine ammonia-lyase | – | 58.055 | Mantilla et al. (111) |
Aquaporin and histidine-ammonia lyase are not present in T. brucei acidocalcisomes.
Transmembrane domain (TMD).
Molecular weight (MW).
Proteomic analysis of C. merolae acidocalcisomes, then called “polyP vacuoles,” identified several orthologs to proteins present in acidocalcisomes of T. brucei, like vacuolar H+-ATPase subunits, vacuolar H+-PPase (CMO102C), endopolyphosphatase (CMG087C), acid phosphatase (CM7279C), zinc transporter (CMF058C), vacuolar iron transporter (CMT466C), and vacuolar transporter chaperone 1 (CMP062C) (63). Antibodies against an M13 family metallopeptidase (CMP249C) or against C-terminal HA-tagged o-methyltransferase (CMT369C), ABC transporter (CMS401C), and prenylated Rab receptor (CMJ260C) co-localized with the vacuolar H+-ATPase to these vacuoles, as detected by immunofluorescence analyses (63).
Early work on C. reinhardtii acidocalcisomes reported the isolation of the then called “polyphosphate bodies” with detection of PPi and polyP using 31P NMR, and phosphorus, magnesium, and calcium, using X-ray microanalysis (58). A PPase activity was also measured in microsomal fractions, and antibodies against the plant V-H+-PPase localized a protein to intermediate-sized vesicles, plasma membranes, and contractile vacuoles (59). Later work that included 45Ca release experiments from acidic compartments, organelle isolation using iodixanol gradient centrifugation, X-ray microanalysis of isolated or in situ organelles, PPi-driven proton transport measurements, and immunofluorescent analysis revealed that these intermediate-sized vesicles were acidocalcisomes and that they possess a V-H+-PPase and a V-H+-ATPase (18). In a recent work, acidocalcisomes were isolated by a modified iodixanol gradient centrifugation method, and proteomic analysis of the fractions obtained was done (86). Two types of vacuoles were obtained and named lysosome-related organelles from stationary phase (stat-LROs) and iron-loaded lysosome related organelles (Fe-LROs) or ACs. The difference between these fractions is that the Fe-LROs were obtained from cells incubated at high iron concentration (>200 µM) for 1 day after partially starving them of iron for 5 days (<50 µM), while the stat-LROs were cultured for 6 days in a regular culture medium (86). Surprisingly, the authors considered them as different organelles. However, the stat-LRO appeared as acidocalcisomes surrounded by portions of cytoplasm [Fig. 1E of reference (86)], contained deformed mitochondria and cellular debris [Fig. S5E and D of reference (86)], and had autophagy markers (ATG8), suggesting that they are autophagosomes. The localization of four proteins [inorganic pyrophosphatase (IPPase, Cre10.g424100.t1.2), copper transporter (CRT1, Cre13.g570600.t1.2), V-H+-ATPase (vATPase subunit C/D, Cre03.g176250.t1.2), and isocitrate lyase (CL1, Cre06.g282800.t1.2)] to acidocalcisomes was validated by their co-localization with DAPI-stained polyP and LysoTracker (86). A number of potential acidocalcisome proteins with similarity to those found in T. brucei acidocalcisomes (metal transporters, phosphate transporters, and Ca2+-ATPases) were found in these fractions (86) and await validation.
In addition to the above studies, several publications have reported the finding of proteins homologous to those in acidocalcisomes of T. brucei (Table 1) and postulated their presence in acidocalcisome-like organelles. The V-H+-PPase was found in acidocalcisomes of Leishmania donovani (47), Plasmodium falciparum (112), P. berghei (20), T. gondii (50), Agrobacterium tumefaciens (39), L. mexicana amazonensis (76), Rhodospirillum rubrum (40), Phytomonas francai (16), and Eimeria tenella (21). A Ca2+-ATPase (51) and a zinc transporter (113) were found in acidocalcisomes of T. gondii. A Cu-ATPase (ATP7) was found near acidocalcisomes of P. berghei or co-localized with the V-H+-PPase to acidocalcisomes and the plant-like vacuole (PLVAC) of T. gondii (114).
We can conclude that the chemical composition of acidocalcisomes is well adapted to their storage function of phosphorus and cations. They have a reduced number of matrix proteins, as compared to other organelles, and several transporters involved in their store function.
ACIDOCALCISOME PUMPS, TRANSPORTERS, ENZYMES, AND CHANNELS
The proton pumps
Acidocalcisomes of some protists (some trypanosomatids, apicomplexan parasites, and algae) possess two electrogenic proton pumps, a V-H+-PPase and a V-H+-ATPase, while other acidocalcisomes possess either a V-H+-PPase or a V-H+-ATPase.
The V-H+-ATPase is a multisubunit complex of about 14 subunits and 2 sectors, the membrane (V0) and the catalytic (V1) sectors. It was originally identified in acidocalcisomes of T. brucei (4) and T. cruzi (3) by the proton pumping and acridine orange staining inhibition by bafilomycin A1 in permeabilized cells. It was also functionally identified in acidocalcisomes of T. evansi (48), T. gondii (19, 115, 116), C. reinhardtii (18), D. discoideum (17), and human platelets (24). Immunofluorescence and/or immunoelectron microscopy studies confirmed the acidocalcisome localization of several subunits in T. brucei (Table 1), T. cruzi (117), T. gondii (118), C. reinhardtii (18), C. merolae (63), and D. discoideum (17). The complex has other subcellular localizations such as the Golgi complex, the plasma membrane, and the endocytic pathway of T. brucei (26), the plasma membrane of T. cruzi (117), the plasma membrane and the PLVAC of T. gondii (116), the plasma membrane and the digestive vacuole of malaria parasites (119–122), and the contractile vacuoles of C. reinhardtii (18) and D. discoideum (17). Downregulation of the expression of some V-H+-ATPase subunits has therefore multiple phenotypic changes that cannot be attributed solely to the disruption of acidocalcisome functions.
The V-H+-PPase is a single subunit H+ pump that uses PPi to establish a H+ gradient and it was first identified in acidocalcisomes of T. cruzi (9). Previously, this pump had been described only in bacteria, archaea, and plants (123). Structural studies of Vigna radiata V-H+-PPase showed that it is present as a homodimer with 16 transmembrane domains (124). The pump is also localized to the Golgi complex and to the plasma membrane of T. cruzi (3, 125), and could be heterologously expressed in yeast, which lacks the pump (126). Its activity is stimulated by K+, and inhibited by Na+ and pyrophosphate analogs, like IDP and AMDP, as occurs with the plant pump (9). As the pump is very abundant in acidocalcisomes, it has been used as a marker for subcellular fractionation of the organelle (10, 26). T. brucei possess two genes expressing very similar proteins (TbVP1 and TbVP2) (127), both localized to acidocalcisomes (14, 89) (Table 1).
Two genes are also expressed in malaria parasites, but the encoded proteins are different. One corresponds to the K+-stimulated type (VP1, type 1) while the other has homology to the plant K+-insensitive type (VP2, type II) (128). The immunofluorescence localization of P. falciparum VP1 in “intracellular bright spots” (112) and “punctate intracellular inclusions” (128), and the localization of P. berghei VP1 in “intracellular vacuoles” (20) using antibodies against the plant pump, provided indirect evidence of a potential acidocalcisome localization. The pump was also localized to the plasma membrane (112, 128) and to the digestive vacuole (120) of P. falciparum. Two genes are also present in the scuticociliate Philasterides decentrarchi, which have homology to genes encoding K+-sensitive (AVP1) and K+-insensitive (AVP2) plant enzymes, the first protein product localized to acidocalcisome-like organelles and the second localized to the alveolar sacs (57).
The evidence for the acidocalcisome localization of the V-H+-PPase in T. gondii is stronger. Immunoelectron microscopy revealed that antibodies against the enzyme co-localized with antibodies against a Ca2+-ATPase to their acidocalcisomes (51). Acidocalcisome fractions were also shown to contain the AMDP-sensitive PPase activity and PPi-driven acridine orange uptake (54). Interestingly, the V-H+-PPase also localizes to the PLVAC of T. gondii (129), where the V-H+-ATPase is also present (118).
There is also strong evidence for the presence of a V-H+-PPase in acidocalcisomes and contractile vacuoles of C. reinhardtii, where it co-localizes with the V-H+-ATPase (18).
It is interesting to note that the presence of a V-H+-ATPase and a V-H+-PPase in the same compartment occurs in several organelles such as the plant vacuole (10, 130), the contractile vacuoles of C. reinardtii (18), and T. cruzi (106), the acidocalcisomes of some trypanosomatids (127), the digestive vacuole of malaria parasites (120), and the PLVAC of T. gondii (129).
The function of these proton pumps in acidocalcisomes is to acidify the organelles and allow secondary transporters to export H+ in exchange of different cations, or to export H+ together with Pi by the Pi symporter (Pho91) present in some species (131). The low pH also contributes to maintain the solubility of polyP. In the case of the V-H+-PPases, they also contribute to maintain low levels of cytosolic PPi and then facilitate anabolic reactions, which produce PPi (132). A third function could be the generation of a membrane potential that could be used to reverse the V-H+-ATPase and generate ATP or reverse the V-H+-PPase to generate PPi, reactions that could occur in vitro (133, 134) but which occurrence in vivo has not been established. In this regard, it has been shown that PPi can generate a membrane potential in acidocalcisomes of T. brucei and T. cruzi that can be reversed by FCCP or CCCP and prevented by AMDP (10, 14).
The P-type ATPases
Two P-type ATPases have been reported in acidocalcisomes of several species, a Ca2+ATPase and a Cu2+-ATPase.
A Ca2+-ATPase activity was initially detected in acidocalcisomes of permeabilized T. brucei procyclic forms by measuring ATP-driven Ca2+ transport inhibited by vanadate (4). The stimulation of H+ uptake (acridine orange uptake) by the Ca2+ chelator EGTA and its release by Ca2+ suggested their accumulation in the same compartment. In addition, H+ uptake was stimulated by vanadate and vanadate inhibited the release of H+ by Ca2+, confirming that the Ca2+-ATPase takes up Ca2+ in exchange for H+ (4). Further work identified a gene encoding an acidocalcisome-located pump that was named PMC1 (for plasma membrane calcium ATPase) because it was closely related to the family of plasma membrane calcium ATPases (PMCA) (96). Its functional role was demonstrated by its ability to complement yeast deficient in the vacuolar Ca2+-ATPase PMC1 (96). The T. brucei pump apparently lacks the C-terminal calmodulin (CaM)-binding domain present in other PMCA pumps (96). Similar Ca2+-ATPases were found in acidocalcisomes of T. cruzi (Tca1) (104), T. gondii (TgA1) (51), and D. discoideum (PAT1) (17, 135), and in vacuoles of S. cerevisiae (PMC1) (136) and E. histolytica (137). Both T. cruzi (104) and T. gondii (52) Ca2+-ATPases were able to complement yeasts deficient in PMC1, providing functional demonstration of their activity. All of them, including PAT1 (135), lack the C-terminal CaM-binding domain although it cannot be ruled out that they could have a non-canonical CaM-binding domain as it has been reported in the T. equiperdum plasma membrane Ca2+-ATPase (138). In this regard, the T. cruzi PMC1 from plasma membrane vesicles was shown to be stimulated by endogenous or mammalian calmodulin (139).
A Cu-ATPase (ATP7) was first detected in the plasma and organellar membranes of T. brucei (94). The intracellular distribution agrees with its acidocalcisome localization by the TrypTag project (89) (Table 1). The N-terminal region of the protein was shown to bind copper in vitro and within E. coli cells (94). Leishmania major ortholog (LmATP7, LmjF33.2090) also localized to vesicles compatible with acidocalcisomes, in addition to the plasma membrane (95). It has been proposed that vesicles containing copper fuse to the plasma membrane to export the metal (140). The essentiality of the CuATPase of T. brucei has not been tested, while complete knockout (KO) of L. major CuATPase was not possible, suggesting its essentiality (95). Interestingly, copper accumulates in acidocalcisomes of C. reinhardtii during zinc limitation, but the potential role of a CuATPase has not been investigated (141). P. berghei and T. gondii CuATPases are also detected in intracellular storage vesicles, which could be acidocalcisomes (114).
PolyP synthesis and degradation
Synthesis and translocation of polyP into the lumen of acidocalcisomes is catalyzed by the vacuolar transporter chaperone (VTC) complex (142). The VTC complex is present in fungi, trypanosomatids, apicomplexans, and algae but is absent in animals. The VTC complex of S. cervisiae has five subunits (Vtc1–Vtc5), of which Vtc4 is the catalytic subunit and forms subcomplexes with Vtc1 and Vtc2 or Vtc3 (Vtc1/Vtc2/Vtc4 or Vtc1/Vtc3/Vtc4), which localize to the vacuole membrane (143). Vtc5 does not form part of these complexes but stimulates their activity. Structural studies of these subcomplexes in yeasts have shown that Vtc1 has three transmembrane domains, while Vtc2, Vtc3, and Vtc4 contain additional SPX (SYG1/Pho81/XPR1) and TTM (triphosphate tunnel metalloenzyme) domains (142, 144, 145). The TTM domain of Vtc4 synthesizes polyP by transferring the γ-phosphate of cytosolic ATP onto the growing polyP chain (142). The SPX domains are receptors for cytosolic inositol pyrophosphates (PP-IPs), which stimulate polyP synthesis (143). PP-IPs bind to the Vtc2 SPX domain, preventing its interaction with the SPX domain of Vtc4 and stimulating its activity (146). The VTC complex has therefore functions of polyP polymerase, polyP translocase, and PP-IPs receptor (144) (Fig. 5).
Fig 5.
A model of the activation mechanism of the VTC complex in yeasts. Schematic of the Vtc4/Vtc3/Vtc1 complex. Subunits are colored. The three subunits of VTC1 are shown in gray; VTC3 is in violet; and VTC4 is in green. PP-InsPs bind to the Vtc3 SPX domain. The star indicates the binding resgion of the SPX domains. ATP is used to add a Pi to the polyP chain at the catalytic domain of VTC4. Key amino acids involved are highlighted. HH, horizontal alpha helix fastening the cytosolic entrance of the transmembrane channel for polyP. (Adapted from reference 145 with permission from the authors.)
Trypanosomatid acidocalcisomes possess orthologs to Vtc1 and Vtc4, which have been studied in T. brucei (91–93), T. cruzi (93), and Leishmania spp. (147, 148). Vtc1 is the smaller subunit of the VTC complex, and the T. brucei protein (TbVtc1) also has three transmembrane domains. The GFP-tagged protein was shown to localize to the acidocalcisomes and endoplasmic reticulum (ER) of T. brucei procyclic forms by immunofluorescence assays (IFAs) and to the acidocalcisomes using polyclonal antibodies against the protein and immunoelectron microscopy (91). Downregulation of TbVtc1 expression by RNAi had marked effects, stopping cell proliferation, changing the morphology and size of acidocalcisomes, and decreasing their V-H+-PPase activity (91). Since the V-H+-PPase generates a protonmotive force that is essential for Ca2+ uptake by the Ca2+/H+ counter-transporting ATPase (4, 14) and for polyP synthesis (13), these decreased V-H+-PPase activities would lead to the decrease in Ca2+ and short- and long-chain polyP content. PolyP deficiency would then result in a deficient response to hypo-osmotic stress (91) and perhaps to cytokinesis defects. However, another possible explanation for the cytokinesis defects could be related to the role of the Vtc1 ortholog in S. pombe (Nrf1), which is an important regulator of the small Rho-like GTPase Cdc42p (149). This GTPase is found in most eukaryotic cells (150). Cdc42p is important for vacuolar function and morphology, critical for cell polarity and cytokinesis and for docking assembly in S. cerevisiae (151, 152).
TbVtc4 also appears to have three transmembrane domains. In situ tagged TbVtc4 (92), as well as tagged T. cruzi Vtc4 (93), localized to acidocalcisomes, as detected by IFA and immunoelectron microscopy. The catalytic domain of either TbVtc4 or TcVtc4 was able to catalyze the synthesis of short-chain polyP (100–300 Pi units), in contrast to the synthesis of long-chain polyP by the catalytic core of ScVtc4 (92, 93). PPi inhibited the synthesis of short-chain polyP by TbVtc4 or TcVtc4 and stimulated the synthesis of polyP by ScVtc4 (92, 93). A conditional KO of TbVtc4 in bloodstream forms resulted in decreased proliferation and short-chain polyP production without affecting long-chain polyP synthesis. These cells had higher sensitivity to hypo-osmotic and hyperosmotic stresses and lower infectivity to animals, but no cytokinesis defects (92). RNAi of TbVtc4 in procyclic forms also reduced proliferation and short-chain polyP synthesis without affecting long-chain polyP synthesis, suggesting that other polyP polymerases might occur in trypanosomes (93). Interestingly, both TbVtc1 and TbVtc4 appear to be palmitoylated (101).
Variable levels of polyP were observed in different Leishmania spp. (148). PolyP was more abundant in late logarithmic growth phase promastigotes of L. major and Leishmania amazonensis and decreased overtime in stationary phase cultures, although protein Vtc4 levels were constant (148). The role of Vtc4 in Leishmania spp. was studied by either knockout or knockdown of its gene expression. PolyP levels were reduced 5- to 10-fold by knockdown of Vtc4 expression by RNAi in Leishmania guyanensis, but this did not affect mice footpad infection (148). L. major Vtc4 knockout promastigotes were devoid of short-chain polyP, grew almost normally in culture, and were able to differentiate normally into metacyclic promastigotes (148). Amastigotes had no detectable Vtc4 and only short-chain polyP (148). While late stationary promastigotes were able to infect macrophages and differentiate into amastigotes, they survived less than wild-type parasites (147, 148). When injected into the footpad of mice, they had a delay in replication, but they were still able to produce lesions (148).
T. gondii has homologs to Vtc2 and Vtc4. A signature-tagged mutagenesis screen identified a disrupted locus that encodes a protein with homology with yeast Vtc2 (153). The tagged TgVtc2 was shown to have a punctate localization, compatible with acidocalcisomes, and the mutant had a significant reduced level of short- and long-chain polyP, which could be partially restored by complementation with an exogenous gene (153). The mutant tachyzoites were able to differentiate into bradyzoites, and they upregulated sixfold the expression of a FIKK kinase (153). Attempts to knockout TgVtc2 were unsuccessful, suggesting that the gene is essential. TgVtc4 was shown to co-localize with the V-H+-ATPase to acidocalcisomes (118), but no further studies were reported.
Evidence for the presence of homologs to Vtc1 and Vtc4 in acidocalcisomes of C. reinhardtii has also been reported (154). A gene encoding a ScVtc1 homolog was found to be deleted in a C. reinhardtii mutant (ars76) with altered ability to acclimate to sulfur deficiency because of its inability to accumulate extracellular or periplasmic arylsulfatase (ARS) (154). The mutant had few acidocalcisomes; was more sensitive to exposure to sulfur-, phosphorus-, or N-deficient conditions; and was defective in trafficking of periplasmic ARS. The phenotypes were complemented by providing an exogenous CrVtc1 gene (154). A truncated CrVtc4 with the putative kinase domain was expressed and shown to synthesize polyP, while polyP was undetectable in a CrVtc4 loss-of-function mutant (155). Phylogenetic studies revealed that VTC and phosphate transporter genes, like C. reinhardtii phosphate transporter C, which transports Pi out of acidocalcisomes, were conserved among species that store phosphorus as vacuolar polyP and absent from genomes of higher plants that store phosphorus as Pi in the vacuoles. This suggests loss of VTC and PTC genes during evolution to higher plants (155). CrVtc1- and CrVtc4-mediated synthesis of polyP was also measured at different times after adding Pi to P-starved cells, and it was found that their expression was markedly reduced after addition of Pi (156). The authors presented indirect evidence that inositol pyrophosphates could be stimulating polyP synthesis. Addition of Pi together with neomycin, which inhibits inositol phosphate synthesis by binding to phosphatidylinositol 4,5-bisphosphate, accumulated less Pi than controls and had less polyP (156).
Degradation of polyP in the acidocalcisome-like vacuole of S. cerevisiae is through the activity of two endopolyphosphatases (they degrade polyP by attacking internal phosphoanhydride bonds): Ppn1 and Ppn2. ScPpn1 was originally described as a homodimer (157) but later found to be a homotetramer of 35-kDa subunits that requires protease activation of a 78-kDa precursor polypeptide (prePpn1) (158). The processed enzyme produces Pi and polyP3 (158). The enzyme requires Mn2+ or Mg2+; is inhibited by Ca2+, Zn2+, Pi, and PPi; and has a neutral pH optimum (7.5) (157). Null mutants in ScPpn1 are growth defective and accumulate long-chain polyP (159). The production of monomeric Pi suggests that ScPpn1 can also display exopolyphosphatase activity (cleavage of the terminal Pi) (158, 160).
ScPpn2 belongs to the phosphoprotein phosphatase family. It has one TMD in the N-terminus and the catalytic domain at its C-terminus, which localizes to the lumen of the acidocalcisome-like vacuole (159–161). ScPpn2 immunoprecipitated from yeast vacuoles had Zn2+- or Co2+-stimulated endopolyphosphatase activity. Studies with different ScPpn1, ScPpn2, and ScVtc mutants concluded that ScPpn1 might be involved in the mobilization of polyP stores, while ScPpn2 might control polyP chain length and that polyP levels of yeasts are controlled through synthesis rather than degradation (161).
Trypanosomatids have an acidocalcisome protein with homology to ScPpn2, but this enzyme has not been studied (Table 1). In addition, trypanosomatids have an acidocalcisome vacuolar soluble pyrophosphatase (VSP), which has exopolyphosphatase activity in the presence of Zn2+ and pyrophosphatase activity in the presence of Mg2+ (99, 109, 162). The enzyme is also present in the cytosol (109). The exopolyphosphatase activity of the T. brucei enzyme is inhibited by bisphosphonates, which also inhibit in vivo infections (163). The enzyme has a putative calcium EF-hand-binding domain that was originally proposed to be involved in its oligomerization (99), but further structural work failed to find evidence for this effect (164). Downregulation of its expression by RNAi in T. brucei bloodstream form (BSF) resulted in decreased levels of short- and long-chain polyP and altered response to phosphate starvation and hypo-osmotic stress (99). Overexpression of the enzyme in T. cruzi led to a decrease in PPi and short- and long-chain polyP, larger acidocalcisomes, defective response to hyperosmotic stress, and lower proliferation in fibroblasts with reduced persistence in tissues of mice (109). The crystal structure of both the T. brucei (164, 165) and T. cruzi (164) enzyme has been solved and revealed an unusual tetrameric oligomeric state containing head-to-tail dimers (Fig. 6).
Fig 6.
Structure superimposition of native TcVSP1 (green) and TbVSP1 (red) in the absence of any added substrate or inhibitor ligands. There is an EF-hand N-terminal domain and a C-terminal PPase domain. A long (~17 residues, ~32A) polypeptide chain linker connects the two domains. Blue denotes divalent cation-binding aspartates. (Reproduced from reference 164 with permission from the American Chemical Society.)
Little is known on the mechanisms involved in polyP degradation in acidocalcisomes of other species except for the presence of a putative endopolyphosphatase in acidocalcisomes of the alga C. merolae (63).
The conclusion is that acidocalcisomes possess mechanisms for the synthesis (VTC complex) and degradation (endo- and exopoylphosphatases) of polyP. The absence of VTC complex in animal cells and the peculiarities of the polyP-degrading enzymes make them potential drug targets against pathogenic organisms possessing these organelles.
Channels
The presence of at least three channels has been reported in acidocalcisomes of trypanosomatids, an aquaporin, a potassium channel, and the inositol 1,4,5-trisphosphate receptor (IP3R) (Tables 1 and 2).
T. cruzi aquaporin 1 (TcAQP1) co-localized with the V-H+-PPase to acidocalcisomes and the contractile vacuole complex (CVC), as revealed by IFA using polyclonal antibodies against the GFP-tagged protein or against a synthetic C-terminal peptide (105). When expressed in Xenopus oocytes, TcAQP1 was shown to be water permeable but not glycerol permeable, in agreement with its phylogenetic analysis that indicates that it belongs to the orthodox (water transporting) aquaporin branch (105). The protein was found to be N-glycosylated (105) and was important for the response of the cells to hypo-osmotic stress (106). Upon hypo-osmotic stress, there was swelling of acidocalcisomes and microtubule- and cyclic AMP-mediated fusion of the organelles to the CVC (Fig. 7) with translocation of TcAQP1 and water release to facilitate cell volume recovery (106). TcAQP1 is also important for the cellular response to hyperosmotic stress as revealed by experiments with inhibitors, or knockdown or overexpressed TcAQP1 (166).
Fig 7.
Contractile vacuole complex of T. cruzi. (A) Epimastigote showing the kinetoplast (K), CV, S, F, and similar electron-dense material present in the CV bladder and in Ac. (B) Three dimensional model of a tomogram showing an acidocalcisome (orange) fusing with the CVC (blue) after hypo-osmotic stress. The CVC is represented by a central vacuole or bladder, and tubules or spongiome. Picture taken by Kildare Miranda. Experimental details in reference (167). (Panel A was reproduced from reference 105 with permission from the American Society for Biochemistry and Molecular Biology.) Ac, acidocalcisome; CV, contractile vacuole; CVC, contractile vacuole complex; F, flagellum; S, spongiome.
A gene with homology to inward rectifying potassium channels was found in T. brucei, and the protein product (TbIRK) was shown to co-localize with V-H+-PPase to acidocalcisomes of procyclic and bloodstream forms (98). The channel was functionally studied by electrophysiology after expressing it in Xenopus oocytes and found to be selective for potassium ions and inhibited by cesium but not by barium (98). The sequence TXTGY(F)G of the selectivity filter found in other potassium channels is replaced by the sequence GGYVG in TbIRK, which was confirmed as the selectivity filter by mutagenesis studies (98). Downregulation of its expression in procyclic forms was obtained by RNAi, but no proliferation changes were found.
Early proteomic studies of contractile vacuole fractions of T. cruzi reported the presence of peptides from a putative IP3R (168). The trypanosomatid IP3Rs possess domains present in other eukaryotes such as the putative suppressor domain-like, ryanodine receptor IP3R homology (RIH), and RIH-associated domains, and a Ca2+-specific selectivity filter, GVGD (169). However, they have five instead of six TMDs and conserve only 4 or 5 of the 10 residues proposed to form a basic pocket that binds IP3 (97). In situ tagging of the ortholog gene in T. brucei co-localized the protein product with the V-H+-PPase to the acidocalcisomes but not to the ER (97). These results were later confirmed using specific antibodies against T. brucei IP3R (TbIP3R) (26, 170) (Fig. 3). The TbIP3R function was confirmed by expressing it in DT40 chicken B lymphocytes that are knockout mutants for the three animal IP3Rs (DT-40–3KO) (171). IP3 was also able to release Ca2+ from permeabilized cells, and isolated acidocalcisomes’ previous acidification of the organelle by PPi. Ca2+ release by uncaging IP3 by UV light was also observed in Fluo4-AM-loaded live procyclic forms (97).
Since acidocalcisomes are rich in phosphorus compounds, the response of TbIP3R expressed in DT-40–3KO cells to the addition of these compounds to its luminal side was investigated using patch-clamp recordings of nuclear membranes, which are continuous with the ER. K+ at 140 nM was the charge carrier, and IP3 was added to the patch pipette (cytosolic side) (172). The currents generated by IP3 were abolished by acidic pH and heparin and inhibited by 2-APB and caffeine (172). Pi and PPi, added to the luminal side, increased the currents, while polyP3, but no longer-chain polyP, inhibited it. These latter effects were not observed when rat IP3R-1 (RnIP3R) was tested instead of TbIP3R (172). In summary, the results suggest that TbIP3R is closed at the acidic conditions of the acidocalcisomes, but alkalinization or polyP hydrolysis favors the channel opening by IP3. Downregulation of TbIP3R expression by RNAi showed that it was essential for proliferation and infectivity of T. brucei in mice (97).
The T. cruzi IP3R was originally proposed to have an ER localization, although no clear co-localization with T. brucei ER marker binding immunoglobulin protein was demonstrated, and the antibodies against the receptor gave a punctate staining in T. cruzi (171). Endogenous tagging of TcIP3R confirmed its acidocalcisome localization (107). Expression of TcIP3R in DT4-3KO cells also established its function in releasing Ca2+ upon stimulation with IP3 (108, 171).
TcIP3R single-allele knockout epimastigotes showed reduced proliferation rate and metacyclogenesis, while trypomastigotes had lower Ca2+ increase upon attachment to host cells and were less infective but did not show differences in amastigote replication (171). TcIP3R double-allele ablation epimastigotes obtained by CRISPR/Cas9 genome editing also had reduced proliferation but increased metacyclogenesis (108). Similar to the single-allele KO, the double-allele KO trypomastigotes are also less infective without differences in amastigote replication. Single knockout trypomastigotes had a significantly higher transformation rate into amastigotes (171). Conversely, although overexpression of TcIP3R in epimastigotes did not modify their proliferation rate, they had reduced metacyclogenesis. Trypomastigotes overexpressing the TcIP3R had increased Ca2+ levels upon cell attachment and higher infectivity in tissue culture cells (171), without affecting (171) or reducing (108) amastigote replication, and showed a decreased transformation rate into amastigotes (171).
Little is known about the presence of channels in acidocalcisomes of other species. A transient receptor potential (TRP)-like protein is in the acidocalcisome-like vacuoles of S. cerevisiae and other fungi (TRPY1 or yvc1) (173). This channel is activated by either mechanical force or Ca2+ and mediates vacuolar Ca2+ release upon hyperosmotic stress (174). The channel has a tetrameric structure displaying activating and inhibiting Ca2+-binding sites and co-purifying with an inhibitory phosphatidylinositol 3-phosphate [PI(3)P] lipid (175). A two-pore channel (TDC2) is localized in the acidocalcisome-like platelet dense granules (PDGs) and regulates PDG luminal pH and functions in Ca2+ release from PDGs forming perigranular Ca2+ microdomains (175).
Membrane transporters
Several metal ion transporters were identified in trypanosomatid acidocalcisomes. Proteomic analysis of T. brucei acidocalcisomes (26) and immunofluorescence analysis of the parasite proteome (89) resulted in the identification of transporters for copper (copper ATPase or TbATP7), zinc (Zn transporters 1 and 2 or TbZnT1 and TbZnT2), and iron or manganese (vacuolar iron transporter or TbVIT) (Table 1).
A zinc transporter (TbZnT1) was identified by proteomic analysis of acidocalcisomes of T. brucei (26). TbZnT1 is a member of the cation diffusion facilitator family (CDF) (176). These transporters function as antiporters of Zn2+, Cd2+, Co2+, and/or Ni2+ with protons. All contain six transmembrane domains and share characteristic motifs, such as a CDF family-specific signature sequence at the C-terminus (177). A second zinc transporter (TbZnT2) was also found in acidocalcisomes (Table 1). A zinc transporter was also described in acidocalcisomes of T. cruzi (Table 2) (103). Zinc has catalyticand structural functions in more than 3,000 human proteins and has also a regulatory function (178). Downregulation of TbZnT1 by RNAi was not lethal in procyclic form or BSF (26), probably because of the presence of other Zn transporters in acidocalcisomes.
A zinc transporter (TgZnT) has also been shown to localize to small vesicles or acidocalcisomes that fuse with the PLVAC in T. gondii tachyzoites (113). The protein has six TMDs and is the sole member of the ZnT family of Zn2+ transporters. TgZnT knockout cells have reduced viability in the presence of extracellular Zn2+ (113).
An ortholog to the vacuolar iron transporter (VIT) originally described in Arabidopsis thaliana (179) and to the yeast Ca2+-sensitive cross-complementer 1 (180) was found in the acidocalcisome proteome of T. brucei (TbVIT1) and co-localized with TbV-H+-PPase (26). These transporters are localized to the plant and yeast vacuole, respectively, and have been involved in iron and manganese sequestration into the vacuoles. VIT family members are not found in humans but are found in other fungi, plants, and in malaria parasites, where it apparently localizes to the ER, and its expression downregulation yields reduced liver and blood infections (181). Knockdown of TbVIT1 by RNAi in both procyclic and bloodstream forms resulted in growth defects with a 44% ± 6% and 41% ± 3% reduction in the number of cells 2 and 4 days after tetracycline addition to bloodstream and procyclic forms trypanosomes, respectively, indicating their essentiality for normal growth (26).
A Pi transporter (TbPho91) with homology to S. cerevisiae Pho91p, which is a low-affinity sodium-phosphate (Na+/Pi) symporter that was proposed to export Pi and Na+ from the vacuole lumen to the cytosol (182), was localized to acidocalcisomes of T. brucei (26, 131). Knockout of TbPho91 expression affected cell proliferation under phosphate starvation and increased the size of acidocalcisomes (131). Expression of TbPho91 in Xenopus laevis oocytes and two-electrode voltage clamp recordings found that 5-diphosphoinositol pentakisphosphate, an inositol pyrophosphate, stimulated sodium-dependent depolarization of the oocyte membrane potential and Pi conductance. This effect depended on the presence of the Pho91 SPX domain (131). Similar requirement for the SPX domain was found when giant vacuoles of yeasts expressing wild type or SPX mutant symporters from T. brucei or yeast were patch-clamped to detect currents (131). The results provided evidence for the role of TbPho91 and Pho91p in Pi and Na+ release from acidocalcisomes to the cytosol and for the role of its SPX domain in mediating its regulation by inositol pyrophosphates.
In contrast with TbPho91, T. cruzi Pho91 localizes mainly to the contractile vacuole complex (183). Overexpression of the gene led to higher levels of polyP while its downregulation reduced the growth rate and polyP levels (183).
Other transporters from T. brucei like a cation transporter, probably involved in Mg2+ uptake, a sulfate transporter, a polyamine transporter, two proteins belonging to the major facilitator superfamily, and a homolog to the multidrug and toxic extrusion protein have not been studied in detail, and their functions are unknown (Table 1).
Other components
There is physiological evidence for the occurrence of Ca2+/H+ and Na+/H+ exchange in T. brucei procyclic trypomastigotes (184, 185) and L. donovani promastigotes (186) acidocalcisomes, but the responsible exchangers have not been identified. Some proteins have been localized to acidocalcisomes of T. brucei but have not been studied in detail, such as the putative enzyme acid phosphatase (26), oxidoreductase, endopolyphosphatase, and palmitoyl acyl transferase 2 (101), and other proteins like FLA-1 like (102) and STRAP (89) (Table 1).
In T. cruzi, the histidine ammonia lyase (HAL) co-localized with the V-H+-PPase to acidocalcisomes, as detected by the epitope tagging the enzyme or with specific antibodies (111). Fusing HAL to the pH sensor pHluorin detected the alkalinization of acidocalcisomes upon addition of histidine, as HAL catalyzes histidine deamination producing ammonia and urocanate (111). The enzyme has five lysine residues in the C-terminal region that are important for its binding to polyP, which inhibits its activity, and for the parasite survival under starvation conditions (111).
ACIDOCALCISOME BIOGENESIS
Acidocalcisomes share with lysosome-related organelles the transport mechanism of membrane proteins. LROs of animal cells were defined as “cell type-specific modifications of the post-Golgi endomembrane system that have a variety of functions and share some common characteristics with lysosomes” (187). Examples include melanosomes, platelet dense granules, basophil granules, and neutrophil azurophil granules (188).
The first evidence that acidocalcisomes were similar to lysosome-related organelles was provided by studies on the role of adaptor protein 3 (AP-3) complex in the transport of membrane proteins to acidocalcisomes of L. major (189). AP complexes mediate vesicular transport of membrane proteins between cellular compartments (190), and AP-3 complex is involved in sorting proteins to lysosome and lysosome-related organelles from the Golgi (191) or endosomes (192). It recognizes its cargo proteins through di-leucine- or tyrosine-based sorting signals (193). The AP-3 complex is a heterotetramer with two large subunits (β3 and δ), one medium subunit (µ) and one small subunit (σ). Deletion of L. major AP-3δ subunit did not apparently affect the multivesicular tubule-lysosome of these parasites but resulted in less acidic acidocalcisomes because of lower V-H+-ATPase and V-H+-PPase content and activity (189). The soluble VSP1 (162), however, was not affected and was detectable in mutant and wild-type cells acidocalcisomes. The AP-3δ knockout promastigotes had a lower proliferation rate and were able to differentiate into metacyclic forms and infect macrophages, although they replicated less intracellularly and had attenuated virulence in mice. All these alterations were complemented in a re-expressing cell line (189). These studies were later confirmed in T. brucei (100). C-terminal tagged T. brucei AP-3β3 and AP-3δ subunits co-localized with V-H+-PPase antibodies to acidocalcisomes, with the Golgi reassembly and stacking protein antibodies to the Golgi and with Rab11 antibodies to recycling endosomes, revealing their dynamic behavior (100). Downregulation of the expression of these adaptins by RNAi resulted in (i) reduction in the proliferation of both procyclic and bloodstream forms; (ii) progressive reduction in the number and calcium and phosphorus content of acidocalcisomes, leading to their disappearance; (iii) lack of detection of the V-H+-PPase by IFA; (iv) a considerable reduction in acidic Ca2+ and in total levels of PPi and polyP; and (v) reduced volume recovery after hypo-osmotic or hyperosmotic stress, compared with control cells (100). As occurred in L. major (189), trafficking of proteins (membrane glycoprotein p67 and luminal cathepsin L) to the lysosome was unaffected, and the mutant parasites were less virulent in mice (100).
Interestingly, mutation of AP-3 subunits in humans causes the Hermansky-Pudlack syndrome, characterized by decreased acidocalcisome-like platelet dense granules, which contain polyP (24), and resulting in bleeding problems (194). Mutations in ALP5 (AP-3δ) or in APS3 (AP-3σ) in S. cerevisiae also resulted in decreased polyP accumulation in the acidocalcisome-like vacuoles associated to mistargeting of the Vtc5 subunit of the VTC complex (195).
In addition of the evidence of the role of AP-3 complex in the transfer of membrane proteins to acidocalcisomes and acidocalcisome-like organelles like platelet dense granules and other LROs (196), the organelles were identified during assembly in the trans-Golgi in a variety of algae and protists (41). Small vesicles were observed fusing with the organelle membrane of C. reinhardtii, C. monoica, and T. brucei, in a manner similar to those of some lysosome-related organelles (197) providing support for their trans-Golgi origin (Fig. 8).
Fig 8.

Acidocalcisomes from C. reinhardtii assembling at the trans-Golgi face. Note the rugose E-face with granule and fusing Golgi vesicles (asterisks); Bar, 100 nm. (Reproduced from reference 41with permission from Elsevier.) g, polyphosphate granule.
Expression downregulation of several proteins has also been reported to affect the biogenesis of acidocalcisomes in trypanosomatids. Null mutants of the first enzyme in the sphingolipid synthesis pathway (serine palmitoyl transferase) of L. major have morphologically altered acidocalcisomes with lower long-chain polyP content (198). Acidocalcisomes of null mutants for target of rapamycin 3 (TOR 3) kinase of L. major are smaller and less numerous with little DAPI staining of polyP, suggesting a lower polyP content (199). In contrast, RNAi of TOR three kinase in T. brucei resulted in larger acidocalcisomes with increased levels of PPi and polyP (200). In addition, RNAi downregulation of a motor kinesin of T. brucei resulted in functional alterations (less Ca2+ release), suggesting that this motor protein could be involved in the traffic of vesicles to acidocalcisomes (201).
In conclusion, beyond the role of the AP-3 complex in the targeting of proteins to the acidocalcisomes, the knowledge is still incomplete because there are probably other trafficking proteins involved in the biogenesis of the organelle. No targeting signals of proteins localized to the acidocalcisomes, besides the presence of one or more tyrosine-based sorting signals with the YXXØ (Ø corresponds to a hydrophobic amino acid) consensus motif (26), have been described.
INTERACTION OF ACIDOCALCISOMES WITH OTHER ORGANELLES
Two types of interactions between acidocalcisomes and other organelles have been reported. The first one is through membrane fusion with the CVC that occurs in T. cruzi (167) and potentially in other species possessing acidocalcisomes and a CVC. The second is through membrane contact sites, which are regions at which organelles interact to exchange biomolecules and that are usually less than 30 nanometers apart (202).
Video microscopy of T. cruzi epimastigotes submitted to hypo-osmotic stress showed vesicle fusion events both with the contractile vacuole and between vesicles (106). GFP-tagged TcAQP1 labeled the acidocalcisomes, but during hypo-osmotic stress, most of the label was translocated to the contractile vacuole (106). It was possible to quantify the translocation by counting the cells with one bright spot corresponding to the CVC since labeling of acidocalcisomes faded with time. Under isosmotic conditions, this translocation of TcAQP1 was stimulated by cyclic AMP analogs and cyclic adenosine monophosphate (cAMP) phosphodiesterase inhibitors, while preincubation with cyclic AMP or microtubule inhibitors decreased translocation under hypo-osmotic stress (106). The results are consistent with microtubule- and cyclic AMP-dependent traffic of acidocalcisomes toward the CVC culminating in fusion of the organelles. Further evidence of fusion of acidocalcisomes and the CVC was provided by the change in localization of GFP-tagged Rab32 or VAMP7 from acidocalcisomes to the CVC under hypo-osmotic stress and electron tomography evidence of fusion of the organelles (167) (Fig. 7B). Acidocalcisomes from C. reinhardtii (18) and D. discoideum (17) also appear to contact their CVCs under hypo-osmotic conditions.
In addition to fusion with the CVC, there is also evidence of fusion of acidocalcisomes between themselves decreasing their numbers and becoming larger, for example, upon hypo-osmotic stress (17, 106) or after knockdown of AP-3 subunits (100) or Vtc1 (91).
Acidocalcisomes of trypanosomatids also interact with other organelles, such as mitochondria, nuclei, and lipid inclusions, through membrane contact sites (75). The interaction with mitochondria has been studied in more detail in T. brucei. In animal cells, the close apposition of the endoplasmic reticulum to the mitochondria facilitates IP3R-dependent Ca2+ transfer, which is important to maintain the mitochondrial bioenergetics (203). A similar situation occurs between acidocalcisomes, where the IP3R localizes in trypanosomatids (97, 107), and the mitochondria, as demonstrated by super-resolution structured illumination microscopy, electron microscopy, proximity ligation assays, and functional studies (202). Super-resolution and electron microscopy revealed that the two organelles appear to be less than 30 nm apart, and proximity ligation assays using antibodies against tagged acidocalcisome TbIP3R and mitochondrial voltage dependent anion channel (VDAC) detected several contacts between the organelles (202). In addition, downregulation of the TcIP3R expression resulted in alterations in the mitochondrial bioenergetics (108).
ROLE IN CALCIUM SIGNALING
Ca2+ signaling is the result of cytosolic Ca2+ increase due to Ca2+ release from intracellular stores or Ca2+ influx through plasma membrane channels and regulates numerous cellular processes including proliferation, differentiation, and cellular motility. In trypanosomatids, acidocalcisomes are major Ca2+ stores and possess mechanisms for Ca2+ uptake (Ca2+-ATPases) and Ca2+ release (IP3R). In other species, acidocalcisomes and acidocalcisome-like structures have been shown to possess Ca2+-ATPases but, except for the acidocalcisome-like vacuole of S. cerevisiae that possesses a Ca2+ release mechanism (Ca2+/H+ exchanger VCX1) (204), it is not known how Ca2+ release occurs and whether it is involved in signaling.
Two important functions have been attributed to the acidocalcisome releasable Ca2+ in T. cruzi. One is the regulation of mitochondrial bioenergetics by the TcIP3R-mediated acidocalcisome Ca2+ release (28) (Fig. 9). Ca2+ is taken up by the mitochondria and results in activation of pyruvate dehydrogenase dephosphorylation and increased O2 consumption. In the absence of TcIP3R, epimastigotes were viable but showed slower proliferation rate and increased metacyclogenesis (108). Trypomastigotes were less infective, and amastigotes did not show changes in replication. In addition to lower O2 consumption in the null mutants, there was an increased phosphorylation of pyruvate dehydrogenase E1a subunit, AMP:ATP ratio, ammonia production, and autophagy (108). Overexpression of TcIP3R led to lower metacyclogenesis, host cell invasion, and amastigote replication (108).
Fig 9.
Acidocalcisomes are in close contact with the mitochondria. Ca2+ (black circles) is released from acidocalcisomes upon stimulation of the IP3 receptor (IP3R) by IP3, and after passing the outer mitochondrial membrane through the highly permeable VDAC, it is handled by the MCU. When in the matrix, Ca2+ stimulates the TCA function and oxidative phosphorylation (OXPHOS) with the generation of ATP, preventing autophagy. MCU, mitochondrial Ca2+ uniporter; VDAC, voltage-dependent anion channel.
Acidocalcisome releasable Ca2+ is also important for host cell invasion by T. cruzi. There is an increase in the parasite cytosolic Ca2+ upon its attachment to host cell that is needed for normal invasion (205, 206). That this Ca2+ increase is coming from the acidocalcisomes is supported by the deficient invasion when the acidocalcisome TcIP3R expression is downregulated (108) or when the phospholipase C, which generates IP3, is inhibited (207). Host cell invasion requires energy (208), and TcIP3R-dependent acidocalcisome Ca2+ release is important to maintain mitochondrial bioenergetics (108). Finally, depletion of acidocalcisome Ca2+ by treatment of the cells with a combination of ionomycin and nigericin inhibits host cell invasion (207).
ROLE IN OSMOREGULATION
This role has been studied mainly in trypanosomatids. Trypanosomatids are submitted to drastic fluctuations in osmolarity during their life cycles. In the insect vector lower digestive tract, epimastigotes of T. cruzi are submitted to increasing osmolarities that reach values of up to 1,000 mosmol/kg in the yellow rectal content (209). When the blood stages of trypanosomatids circulate through the kidney of their mammalian hosts, they need to resist up to 1,200–1,400 mosmol/kg in the ascending limb of the vasa recta and return to isosmotic conditions of 300 mosmol/kg a few seconds later (210). In addition, these parasites, like all cells, need to regulate their volumes continuously (28).
As described above (Interaction of Acidocalcisomes with Other Organelles), upon hypo-osmotic stress, acidocalcisomes of T. cruzi take part in a complex pathway that leads to their fusion with the contractile vacuole complex and transfer of TcAQP1 (106). A model was proposed (106) suggesting that cell swelling, by activating a mechanosensitive channel, causes an increase in cAMP resulting in a microtubule-dependent movement and fusion of acidocalcisomes with the contractile vacuole and translocation of TcAQP1. This fusion would also lead to transfer of osmolytes (cations and phosphorus resulting from the hydrolysis of polyP) to the CVC, leading to water uptake that could then be released out of the cells. The process would be terminated by a phosphodiesterase that hydrolyzes cAMP to 5′-AMP. Some evidence in favor of this model is that (i) a mechanosensitive channel (211), adenylyl cyclases (212), and a cAMP phosphodiesterase C (213) are present in the CVC; (ii) fusion of acidocalcisomes to the CVC is enhanced by cAMP analogs and PDE inhibitors and inhibited by microtubule and adenylyl cyclase inhibitors (106); and (iii) there is hydrolysis of polyP upon hypo-osmotic stress (13) (Fig. 10).
Fig 10.
Model proposed for regulatory volume decrease in T. cruzi. Cell swelling causes activation of an adenylyl cyclase, probably by a mechanosensitive channel. cAMP formed stimulates the microtubule-dependent fusion of acidocalcisomes with the contractile vacuole and translocation of an aquaporin. A rise in ammonia in acidocalcisomes activates an exopolyphosphatase activity, which cleaves polyP, releasing inorganic phosphate residues and polyP-chelated cations. The resulting osmotic gradient sequesters water through the aid of the aquaporin. Water is released into the flagellar pocket. Amino acid release contributes to the volume recovery. A, acidocalcisome, AA, amino acid; CV, contractile vacuole.
The involvement of acidocalcisomes in the response to hypo-osmotic stress has also been reported in other trypanosomatids. L. major acidocalcisomes were reported to lose Na+ and Cl− (214), and T. brucei acidocalcisome V-H+-PPase downregulation led to a deficient regulatory volume decrease after hypo-osmotic stress (127).
On the other hand, hyperosmotic stress results in increased synthesis of acidocalcisome polyP (13), which could have a role in sequestering inorganic ions. This sequestration would reduce the ionic strength increased by water elimination and prevent cell damage (166, 215).
ROLE IN CATION AND PHOSPHORUS STORAGE
Acidocalcisomes store phosphorous compounds such as Pi, PPi, and polyP, combined with inorganic (Ca2+, Mg2+, Zn2+, Fe2+, Cu2+, Mn2+, K+, and Na+) and organic cations, like lysine, arginine, ornithine, and polyamines. Except for the acidocalcisome-like platelet dense granules (24), or the secretory granules of mast cells (25), which can release the intact polyP polymer, phosphorus release from most acidocalcisomes is in the form of orthophosphate and through Pi transporters like TbPho91 and S. cerevisiae Pho91p (131). Dictyostelium discoideum also releases polyP (~9 mer) (216), but, although their acidocalcisome-like mass dense granules are rich in polyP (17), it is not known whether they are the source of extracellular polyP.
PolyP release in the case of human platelets has been extensively studied since it was found to have roles in blood clotting and fibrinolysis (217). Platelet-size polyP (~60 to 100 mer) accelerates blood clotting by activating the contact pathway, promoting factor V activation, and inhibiting the function of tissue factor pathway inhibitor (TFTPI). It was found that platelet-size polyP accelerates factor V activation by factor Xa, thrombin (217, 218), and factor XIa (219) and promotes factor XI back-activation by thrombin (220). It also delays clot lysis by promoting the thrombin-activatable fibrinolysis inhibitor (217) (Fig. 11). The ability of polyP to activate the contact pathway was shown to depend on the polymer length being more relevant with long-chain polyP, like that present in bacteria (218). Long-chain polyP found in bacteria is a potent proinflammatory agent (221), suppresses complement (222), and modulates fibrin clot structure and stability (223). PolyP secreted from mast cells could be important for their proinflammatory and procoagulant activity (25).
Fig 11.
Steps of the coagulation cascade affected by polyP. PolyP (red) accelerates factor V activation by factors Xa and thrombin, accelerates factor IX back-activation by thrombin, inhibits the ability of TFPI to inhibit factor Xa, and enhances fibrin polymerization. LC, long-chain polyP; polyP, polyphosphate.
Storage of polyP in acidocalcisomes could be important as a phosphorus reserve for the synthesis of nucleic acids and phospholipids when the cells multiply. T. cruzi epimastigotes increase their synthesis of short- and long-chain polyP during the lag phase of proliferation and then consume it when they start dividing (13). PolyP is also higher during the logarithmic phase of growth and decreases in Leishmania spp. during stationary phase (148). There is also an increase in polyP synthesis when trypomastigotes differentiate into amastigotes, which are the forms that multiply within the host cells and possibly need to store polyP for their rapid intracellular replication (13). PolyP hydrolysis and synthesis also occurs under hypo-osmotic and hyperosmotic stress, respectively, and phosphorus and cations could serve as osmolytes with a role in cell recovery after osmotic stress (215). Finally, polyP could serve as chelating agent for inorganic and organic cations allowing their safe storage within acidocalcisomes. Some of these cations are essential for cell growth as they are required for a variety of reactions. For example, in trypanosomes zinc is a catalytic/structural cofactor of many metalloproteins; copper is a cofactor for Cu-containing enzymes, such as cytochrome c oxidase; iron is a cofactor of important enzymes, like iron superoxide dismutases (224); and manganese is required for protein glycosylation in the Golgi complex (225). These nutrient metal ions are concentrated for eventual delivery to the cytosol or other organelles. However, these metals are deleterious if they achieve high concentrations in the cell cytosol. In some cases, like with iron and copper, these metals can generate reactive oxygen species via Fenton-based reactions (226). In other cases, like with zinc and manganese, metabolic enzymes can be inhibited (227). Sequestration of these metals within acidocalcisomes could prevent their deleterious cellular reactions. This sequestration is driven by the pH gradient with the cytosol as many of these transporters take up metals in exchange for protons, which are taken up by the vacuolar H+-ATPase (V-ATPase) or the vacuolar H+-pyrophosphatase (VP1) localized to acidocalcisomes. The complexes of these metals with polyP within acidocalcisomes would help to prevent their release.
Except for Ca2+, for which there is a specialized channel, little is known on how other inorganic and organic cations are released from acidocalcisomes.
OTHER ROLES OF ACIDOCALCISOMES
Other roles of acidocalcisomes include their involvement in autophagy, regulation of pH homeostasis, and infectivity.
The finding that blocking acidocalcisome biogenesis in T. brucei by downregulation of the expression of δ and β3 subunits of the AP-3 complex by RNAi inhibited amino acid starvation-induced autophagy revealed the involvement of acidocalcisomes in the initiation of the autophagic process (228). There was a correlation between acidocalcisome acidification and autophagy. Acidocalcisome acidification increased when autophagy was induced by starvation or vanadate treatment, while acidocalcisome alkalinization by bafilomycin A1, monensin, or ionomycin and NH4Cl treatment inhibited autophagy (228). Notably, RNAi depletion of neither the V-H+-ATPase nor the V-H+-PPase, which acidify the acidocalcisomes, showed inhibition of autophagy. Phosphatidylinositol 3-phosphate associates with acidocalcisomes upon starvation and could be involved in autophagy initiation (228).
It was found that when the expression of the V-H+-PPase of T. brucei procyclic forms was downregulated, the cells recovered their intracellular pH after acidification at a slower rate than controls and to a more acidic final pH, suggesting the involvement of these acidic compartment in pH homeostasis (127).
There is indirect evidence that acidocalcisome polyP is important for infectivity. Expression downregulation of acidocalcisome enzymes like the V-H+-PPase (127) or Vtc4 (92), the catalytic subunit of the VTC complex in T. brucei, decreased polyP levels and virulence in mice. Expression downregulation of other proteins involved in the biogenesis of acidocalcisomes such as the AP-3 complex subunits in T. brucei (100) and the TOR kinase three in L. major (199) also reduces parasite virulence. Similarly, overexpression of the acidocalcisome-located VSP in T. cruzi (109) makes the parasite less prone to cause a persistent mouse infection.
ACIDOCALCISOMES AS DRUG TARGETS
The presence and relevance of acidocalcisomes in pathogenic eukaryotes suggest that they can be drug targets (229). Several acidocalcisome components are present in pathogens but absent in animals, like the V-H+-PPase (9, 14, 20, 50, 115, 120, 127, 128), VTC complex (91–93, 148, 153), Pho91 (131), and VIT (26), while others are distantly related to those present in animals, like VSP (99, 163, 164), V-H+-ATPase (117, 118), and Ca2+-ATPase (96, 104). Some of these pumps and enzymes have been screened for potential inhibitors (230). For example, the V-H+-PPase from plants is inhibited by the pyrophosphate analogs known as bisphosphonates, which contain a non-hydolizable P-C-P instead of a P-O-P backbone (231). AMDP and IDP, which have a non-hydrolyzable P-N-P group, are currently used as in vitro inhibitors of the enzyme (232). AMDP was shown to inhibit intracellular replication of T. gondii without affecting host infection (50, 233). The VSP present in trypanosomatids (99, 162) is essential for their normal proliferation and is inhibited by bisphosphonates (163), some of which are active in mice infections by T. brucei (99). Some drugs, like diamidines (234), accumulate in acidocalcisomes, although the involvement of this phenomenon in their mode of action is not known.
In summary, there is convincing evidence that acidocalcisomes are potential drug targets and that continuing their study could contribute to new approaches for chemotherapy of protozoan parasite infections.
CONCLUSIONS AND FUTURE DIRECTIONS
Significant progress has been made in understanding the structure and function of acidocalcisomes in several species. Acidocalcisomes and acidocalcisome-like organelles have been identified in most phylogenetic groups. The search for acidocalcisome-like organelles led to the unexpected finding of polyP in human platelet dense granules (24) and mast cell granules (25) and its secretion, which led to the discovery of the role of this polymer in blood coagulation (217), thrombosis, and inflammation (235). Electron microscopy studies have been fundamental for understanding their ultrastructure (41) and the elemental organization within the organelle (78). Progress has been made in the characterization of the channels, pumps, transporters, and enzymes of these organelles (26) and in their chemical composition (88).These studies led to the discovery in eukaryotes of enzymes previously found only in plants and archaea (V-H+-PPase) (9) and to the first eukaryotic enzymes involved in the synthesis of polyP (VTC complex) (142). Chlamydomonas acidocalcisomes have been developed as models for the study of organellar metal acquisition (85). Trypanosomatids and yeasts have been especially useful for advancing our understanding of the function of their different components and their involvement in cell signaling. The regulation of polyP synthesis in the acidocalcisome-like organelles of yeasts (143) and of the Pi release from acidocalcisomes of trypanosomes (131) by inositol pyrophosphates was first investigated. The finding of mechanisms for Ca2+ uptake and release in acidocalcisomes resulted in the discovery of the role of these organelles rather than the endoplasmic reticulum in the regulation of mitochondrial bioenergetics in trypanosomatids (108). A novel role in osmoregulation involving fusion of acidocalcisomes with the contractile vacuole of T. cruzi (106), potentially present in other species possessing these organelles, was described. Future studies should include the investigation of (i) the role of the enzymes and transporters newly identified; (ii) the reason for organic cation accumulation and whether polyamines are present in all acidocalcisomes; (iii) the tethers that link acidocalcisomes to other organelles; (iv) whether the membrane contact sites are used for transfer of polyP or cations between organelles; and (v) additional roles of acidocalcisomes in cell signaling. In addition, it would be interesting to investigate whether some acidocalcisome-like vacuoles present in ciliates, other lysosome-related organelles of animal species, or organelles accumulating metals, such as the zincosomes (236), possess polyP.
ACKNOWLEDGMENTS
We thank Richard J. Wheeler (University of Oxford) for providing the list of putative acidocalcisome proteins identified by the TrypTag project in Trypanosoma brucei and Ursula Goodenough (Washington University) for useful comments.
Cited work from this laboratory and preparation of this review were supported by grant AI173402 from the National Institute of Allergy and Infectious Diseases. R.D. is the Barbara and Sanford Orkin/Georgia Research Alliance Eminent Scholar.
Biography

Roberto Docampo received an MD in 1972 from the University of Buenos Aires, a PhD in Microbiology in 1977 from the Federal University of Rio de Janeiro (UFRJ), and a PhD in Biological Chemistry in 1979 from the University of Buenos Aires. He then spent four years as a Visiting Scientist at the NIH before joining first, the UFRJ, and then the Rockefeller University, as Visiting Professor. He joined the University of Illinois at Urbana-Champaign in 1990 and in 2005 he moved to the University of Georgia where he remains as Distinguished Research Professor of Cellular Biology and Eminent Scholar. Docampo is best known for his work on the discovery and characterization of acidocalcisomes of different species. His lab was the first to determine the presence and secretion of polyphosphate by human platelets, which led to the discovery of its role in blood clotting. Docampo is Fellow of the American Academy of Microbiology and of the American Association for the Advancement of Science.
Contributor Information
Roberto Docampo, Email: rdocampo@uga.edu.
Corrella S. Detweiler, University of Colorado Boulder, Boulder, Colorado, USA
REFERENCES
- 1. Babes V. 1895. Beobachtungen über die metachromatischen köperchen, sporenbildung, verzweigung, kolben- und kapsel-bildung pathogener bakterien. Zentrlbl Bakteriol Parasitenkd Infektioskr Hyg 20:412–420. [Google Scholar]
- 2. Dell’Angelica EC. 2009. AP-3-dependent trafficking and disease: the first decade. Curr Opin Cell Biol 21:552–559. doi: 10.1016/j.ceb.2009.04.014 [DOI] [PubMed] [Google Scholar]
- 3. Docampo R, Scott DA, Vercesi AE, Moreno SNJ. 1995. Intracellular Ca2+ storage in acidocalcisomes of Trypanosoma cruzi. Biochem J 310 (Pt 3):1005–1012. doi: 10.1042/bj3101005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Vercesi AE, Moreno SN, Docampo R. 1994. Ca2+/H+ exchange in acidic vacuoles of Trypanosoma brucei. Biochem J 304 (Pt 1):227–233. doi: 10.1042/bj3040227 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Scott DA, Docampo R, Dvorak JA, Shi S, Leapman RD. 1997. In situ compositional analysis of acidocalcisomes in Trypanosoma cruzi. J Biol Chem 272:28020–28029. doi: 10.1074/jbc.272.44.28020 [DOI] [PubMed] [Google Scholar]
- 6. LeFurgey A, Ingram P, Blum JJ. 1990. Elemental composition of polyphosphate-containing vacuoles and cytoplasm of Leishmania major. Mol Biochem Parasitol 40:77–86. doi: 10.1016/0166-6851(90)90081-v [DOI] [PubMed] [Google Scholar]
- 7. Dvorak JA, Engel JC, Leapman RD, Swyt CR, Pella PA. 1988. Trypanosoma cruzi: elemental composition heterogeneity of cloned stocks. Mol Biochem Parasitol 31:19–26. doi: 10.1016/0166-6851(88)90141-7 [DOI] [PubMed] [Google Scholar]
- 8. Vickerman K, Tetley L. 1977. Recent ultrastructural studies on trypanosomes. Ann Soc Belg Med Trop 57:441–457. [PubMed] [Google Scholar]
- 9. Scott DA, de Souza W, Benchimol M, Zhong L, Lu HG, Moreno SN, Docampo R. 1998. Presence of a plant-like proton-pumping pyrophosphatase in acidocalcisomes of Trypanosoma cruzi. J Biol Chem 273:22151–22158. doi: 10.1074/jbc.273.34.22151 [DOI] [PubMed] [Google Scholar]
- 10. Scott DA, Docampo R. 2000. Characterization of isolated acidocalcisomes of Trypanosoma cruzi. J Biol Chem 275:24215–24221. doi: 10.1074/jbc.M002454200 [DOI] [PubMed] [Google Scholar]
- 11. Moreno B, Urbina JA, Oldfield E, Bailey BN, Rodrigues CO, Docampo R. 2000. 31P NMR spectroscopy of Trypanosoma brucei, Trypanosoma cruzi, and Leishmania major. Evidence for high levels of condensed inorganic phosphates. J Biol Chem 275:28356–28362. doi: 10.1074/jbc.M003893200 [DOI] [PubMed] [Google Scholar]
- 12. Moreno Benjamin, Rodrigues CO, Bailey BN, Urbina JA, Moreno SNJ, Docampo R, Oldfield E. 2002. Magic-angle spinning 31P NMR spectroscopy of condensed phosphates in parasitic protozoa: visualizing the invisible. FEBS Lett 523:207–212. doi: 10.1016/s0014-5793(02)02977-0 [DOI] [PubMed] [Google Scholar]
- 13. Ruiz FA, Rodrigues CO, Docampo R. 2001. Rapid changes in polyphosphate content within acidocalcisomes in response to cell growth, differentiation, and environmental stress in Trypanosoma cruzi. J Biol Chem 276:26114–26121. doi: 10.1074/jbc.M102402200 [DOI] [PubMed] [Google Scholar]
- 14. Rodrigues CO, Scott DA, Docampo R. 1999. Characterization of a vacuolar pyrophosphatase in Trypanosoma brucei and its localization to acidocalcisomes. Mol Cell Biol 19:7712–7723. doi: 10.1128/MCB.19.11.7712 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Rodrigues CO, Ruiz FA, Vieira M, Hill JE, Docampo R. 2002. An acidocalcisomal exopolyphosphatase from Leishmania major with high affinity for short chain polyphosphate. J Biol Chem 277:50899–50906. doi: 10.1074/jbc.M208940200 [DOI] [PubMed] [Google Scholar]
- 16. Miranda K, Rodrigues CO, Hentchel J, Vercesi A, Plattner H, de Souza W, Docampo R. 2004. Acidocalcisomes of Phytomonas francai possess distinct morphological characteristics and contain iron. Microsc Microanal 10:647–655. doi: 10.1017/S1431927604040887 [DOI] [PubMed] [Google Scholar]
- 17. Marchesini N, Ruiz FA, Vieira M, Docampo R. 2002. Acidocalcisomes are functionally linked to the contractile vacuole of Dictyostelium discoideum. J Biol Chem 277:8146–8153. doi: 10.1074/jbc.M111130200 [DOI] [PubMed] [Google Scholar]
- 18. Ruiz FA, Marchesini N, Seufferheld M, Docampo R. 2001. The polyphosphate bodies of Chlamydomonas reinhardtii possess a proton-pumping pyrophosphatase and are similar to acidocalcisomes. J Biol Chem 276:46196–46203. doi: 10.1074/jbc.M105268200 [DOI] [PubMed] [Google Scholar]
- 19. Moreno SN, Zhong L. 1996. Acidocalcisomes in Toxoplasma gondii tachyzoites. Biochem J 313 (Pt 2):655–659. doi: 10.1042/bj3130655 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Marchesini N, Luo S, Rodrigues CO, Moreno SN, Docampo R. 2000. Acidocalcisomes and a vacuolar H+-pyrophosphatase in malaria parasites. Biochem J 347 Pt 1:243–253. [PMC free article] [PubMed] [Google Scholar]
- 21. Soares Medeiros LC, Gomes F, Maciel LRM, Seabra SH, Docampo R, Moreno S, Plattner H, Hentschel J, Kawazoe U, Barrabin H, de Souza W, Damatta RA, Miranda K. 2011. Volutin granules of Eimeria parasites are acidic compartments and have physiological and structural characteristics similar to acidocalcisomes. J Eukaryot Microbiol 58:416–423. doi: 10.1111/j.1550-7408.2011.00565.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Ramos IB, Miranda K, Pace DA, Verbist KC, Lin F-Y, Zhang Y, Oldfield E, Machado EA, De Souza W, Docampo R. 2010. Calcium- and polyphosphate-containing acidic granules of sea urchin eggs are similar to acidocalcisomes, but are not the targets for NAADP. Biochem J 429:485–495. doi: 10.1042/BJ20091956 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Motta LS, Ramos IB, Gomes FM, de Souza W, Champagne DE, Santiago MF, Docampo R, Miranda K, Machado EA. 2009. Proton-pyrophosphatase and polyphosphate in acidocalcisome-like vesicles from oocytes and eggs of Periplaneta americana. Insect Biochem Mol Biol 39:198–206. doi: 10.1016/j.ibmb.2008.11.003 [DOI] [PubMed] [Google Scholar]
- 24. Ruiz FA, Lea CR, Oldfield E, Docampo R. 2004. Human platelet dense granules contain polyphosphate and are similar to acidocalcisomes of bacteria and unicellular eukaryotes. J Biol Chem 279:44250–44257. doi: 10.1074/jbc.M406261200 [DOI] [PubMed] [Google Scholar]
- 25. Moreno-Sanchez D, Hernandez-Ruiz L, Ruiz FA, Docampo R. 2012. Polyphosphate is a novel pro-inflammatory regulator of mast cells and is located in acidocalcisomes. J Biol Chem 287:28435–28444. doi: 10.1074/jbc.M112.385823 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Huang G, Ulrich PN, Storey M, Johnson D, Tischer J, Tovar JA, Moreno SNJ, Orlando R, Docampo R. 2014. Proteomic analysis of the acidocalcisome, an organelle conserved from bacteria to human cells. PLoS Pathog 10:e1004555. doi: 10.1371/journal.ppat.1004555 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Patel S, Docampo R. 2010. Acidic calcium stores open for business: expanding the potential for intracellular Ca2+ signaling. Trends Cell Biol 20:277–286. doi: 10.1016/j.tcb.2010.02.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Docampo R, Huang G. 2022. New insights into the role of acidocalcisomes in trypanosomatids. J Eukaryotic Microbiology 69. doi: 10.1111/jeu.12899 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Meyer A. 1904. Orientierende untersuchungen über verbreitung. morphologie, und chemie des volutin. Bot Zeit 62:113–152. [Google Scholar]
- 30. Wiame JM. 1947. Étude d'une substance polyphosphorée, basophile et métachromatique chez les levures. Biochimica et Biophysica Acta 1:234–255. doi: 10.1016/0006-3002(47)90137-6 [DOI] [Google Scholar]
- 31. Allan RA, Miller JJ. 1980. Influence of S-adenosylmethionine on DAPI-induced fluorescence of polyphosphate in the yeast vacuole. Can J Microbiol 26:912–920. doi: 10.1139/m80-158 [DOI] [PubMed] [Google Scholar]
- 32. Swellengrebel NH. 1908. La volutine chez les trypanosomes. C R Soc Biol Paris 64:38–43. [Google Scholar]
- 33. Remonsellez F, Orell A, Jerez CA. 2006. Copper tolerance of the thermoacidophilic archaeon Sulfolobus metallicus: possible role of polyphosphate metabolism. Microbiology (Reading) 152:59–66. doi: 10.1099/mic.0.28241-0 [DOI] [PubMed] [Google Scholar]
- 34. Toso DB, Henstra AM, Gunsalus RP, Zhou ZH. 2011. Structural, mass and elemental analyses of storage granules in methanogenic archaeal cells. Environ Microbiol 13:2587–2599. doi: 10.1111/j.1462-2920.2011.02531.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Friedberg I, Avigad G. 1968. Structures containing polyphosphate in Micrococcus lysodeikticus. J Bacteriol 96:544–553. doi: 10.1128/jb.96.2.544-553.1968 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Racki LR, Tocheva EI, Dieterle MG, Sullivan MC, Jensen GJ, Newman DK. 2017. Polyphosphate granule biogenesis is temporally and functionally tied to cell cycle exit during starvation in Pseudomonas aeruginosa. Proc Natl Acad Sci U S A 114:E2440–E2449. doi: 10.1073/pnas.1615575114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Gomes E, Shorter J. 2019. The molecular language of membraneless organelles. J Biol Chem 294:7115–7127. doi: 10.1074/jbc.TM118.001192 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Wang X, Shi C, Mo J, Xu Y, Wei W, Zhao J. 2020. An inorganic biopolymer polyphosphate controls positively charged protein phase transitions. Angew Chem Int Ed 59:2679–2683. doi: 10.1002/anie.201913833 [DOI] [PubMed] [Google Scholar]
- 39. Seufferheld M, Vieira MCF, Ruiz FA, Rodrigues CO, Moreno SNJ, Docampo R. 2003. Identification of organelles in bacteria similar to acidocalcisomes of unicellular eukaryotes. J Biol Chem 278:29971–29978. doi: 10.1074/jbc.M304548200 [DOI] [PubMed] [Google Scholar]
- 40. Seufferheld M, Lea CR, Vieira M, Oldfield E, Docampo R. 2004. The H+-pyrophosphatase of Rhodospirillum rubrum is predominantly located in polyphosphate-rich acidocalcisomes. J Biol Chem 279:51193–51202. doi: 10.1074/jbc.M406099200 [DOI] [PubMed] [Google Scholar]
- 41. Goodenough U, Heiss AA, Roth R, Rusch J, Lee JH. 2019. Acidocalcisomes: ultrastructure, biogenesis, and distribution in microbial Eukaryotes. Protist 170:287–313. doi: 10.1016/j.protis.2019.05.001 [DOI] [PubMed] [Google Scholar]
- 42. Frank C, Jendrossek D, Stabb EV. 2020. Acidocalcisomes and polyphosphate granules are different subcellular structures in Agrobacterium tumefaciens . Appl Environ Microbiol 86. doi: 10.1128/AEM.02759-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Aschar-Sobbi R, Abramov AY, Diao C, Kargacin ME, Kargacin GJ, French RJ, Pavlov E. 2008. High sensitivity, quantitative measurements of polyphosphate using a new DAPI-based approach. J Fluoresc 18:859–866. doi: 10.1007/s10895-008-0315-4 [DOI] [PubMed] [Google Scholar]
- 44. Diaz JM, Ingall ED. 2010. Fluorometric quantification of natural inorganic polyphosphate. Environ Sci Technol 44:4665–4671. doi: 10.1021/es100191h [DOI] [PubMed] [Google Scholar]
- 45. Brock J, Rhiel E, Beutler M, Salman V, Schulz-Vogt HN. 2012. Unusual polyphosphate inclusions observed in a marine Beggiatoa strain. Antonie Van Leeuwenhoek 101:347–357. doi: 10.1007/s10482-011-9640-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Lu HG, Zhong L, Chang KP, Docampo R. 1997. Intracellular Ca2+ pool content and signaling and expression of a calcium pump are linked to virulence in Leishmania mexicana amazonesis amastigotes. J Biol Chem 272:9464–9473. doi: 10.1074/jbc.272.14.9464 [DOI] [PubMed] [Google Scholar]
- 47. Rodrigues CO, Scott DA, Docampo R. 1999. Presence of a vacuolar H+-pyrophosphatase in promastigotes of Leishmania donovani and its localization to a different compartment from the vacuolar H+-ATPase. Biochem J 340 (Pt 3):759–766. [PMC free article] [PubMed] [Google Scholar]
- 48. Mendoza M, Mijares A, Rojas H, Rodríguez JP, Urbina JA, DiPolo R. 2002. Physiological and morphological evidences for the presence acidocalcisomes in Trypanosoma evansi: single cell fluorescence and 31P NMR studies. Mol Biochem Parasitol 125:23–33. doi: 10.1016/s0166-6851(02)00166-4 [DOI] [PubMed] [Google Scholar]
- 49. Miranda K, Docampo R, Grillo O, de Souza W. 2004. Acidocalcisomes of trypanosomatids have species-specific elemental composition. Protist 155:395–405. doi: 10.1078/1434461042650361 [DOI] [PubMed] [Google Scholar]
- 50. Rodrigues CO, Scott DA, Bailey BN, De Souza W, Benchimol M, Moreno B, Urbina JA, Oldfield E, Moreno SN. 2000. Vacuolar proton pyrophosphatase activity and pyrophosphate (PPi) in Toxoplasma gondii as possible chemotherapeutic targets. Biochem J 349 Pt 3:737–745. doi: 10.1042/bj3490737 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Luo S, Vieira M, Graves J, Zhong L, Moreno SN. 2001. A plasma membrane-type Ca2+-ATPase co-localizes with a vacuolar H+-pyrophosphatase to acidocalcisomes of Toxoplasma gondii. EMBO J 20:55–64. doi: 10.1093/emboj/20.1.55 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Luo S, Ruiz FA, Moreno SNJ. 2005. The acidocalcisome Ca2+-ATPase (TgA1) of Toxoplasma gondii is required for polyphosphate storage, intracellular calcium homeostasis and virulence. Mol Microbiol 55:1034–1045. doi: 10.1111/j.1365-2958.2004.04464.x [DOI] [PubMed] [Google Scholar]
- 53. Rodrigues CO, Ruiz FA, Rohloff P, Scott DA, Moreno SNJ. 2002. Characterization of isolated acidocalcisomes from Toxoplasma gondii tachyzoites reveals a novel pool of hydrolyzable polyphosphate. J Biol Chem 277:48650–48656. doi: 10.1074/jbc.M208990200 [DOI] [PubMed] [Google Scholar]
- 54. Rohloff P, Miranda K, Rodrigues JCF, Fang J, Galizzi M, Plattner H, Hentschel J, Moreno SNJ, Langsley G. 2011. Calcium uptake and proton transport by acidocalcisomes of Toxoplasma gondii . PLoS ONE 6:e18390. doi: 10.1371/journal.pone.0018390 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Ruiz FA, Luo S, Moreno SNJ, Docampo R. 2004. Polyphosphate content and fine structure of acidocalcisomes of Plasmodium falciparum. Microsc Microanal 10:563–567. doi: 10.1017/S1431927604040875 [DOI] [PubMed] [Google Scholar]
- 56. Mallo N, Lamas J, de Felipe A-P, Sueiro R-A, Fontenla F, Leiro J-M. 2016. Role of H+-pyrophosphatase activity in the regulation of intracellular pH in a scuticociliate parasite of turbot: physiological effects. Exp Parasitol 169:59–68. doi: 10.1016/j.exppara.2016.07.012 [DOI] [PubMed] [Google Scholar]
- 57. Folgueira I, Lamas J, Sueiro RA, Leiro JM. 2021. Molecular characterization and transcriptional regulation of two types of H+-pyrophosphatases in the scuticociliate parasite Philasterides dicentrarchi. Sci Rep 11:8519. doi: 10.1038/s41598-021-88102-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Komine Y, Eggink LL, Park H, Hoober JK. 2000. Vacuolar granules in Chlamydomonas reinhardtii: polyphosphate and a 70-kDa polypeptide as major components. Planta 210:897–905. doi: 10.1007/s004250050695 [DOI] [PubMed] [Google Scholar]
- 59. Robinson DG, Hoppenrath M, Oberbeck K, Luykx P, Ratajczak R. 1998. Localization of pyrophosphatase and V-ATPase in Chlamydomonas reinhardtii. Botanica Acta 111:108–122. doi: 10.1111/j.1438-8677.1998.tb00685.x [DOI] [Google Scholar]
- 60. Zhu J, Loubéry S, Broger L, Zhang Y, Lorenzo‐Orts L, Utz‐Pugin A, Fernie AR, Young‐Tae C, Hothorn M. 2020. A genetically validated approach for detecting inorganic polyphosphates in plants. The Plant Journal 102:507–516. doi: 10.1111/tpj.14642 [DOI] [PubMed] [Google Scholar]
- 61. Pick U, Weiss M. 1991. Polyphosphate hydrolysis within acidic vacuoles in response to amine-induced alkaline stress in the halotolerant alga Dunaliella salina. Plant Physiol 97:1234–1240. doi: 10.1104/pp.97.3.1234 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Shebanova A, Ismagulova T, Solovchenko A, Baulina O, Lobakova E, Ivanova A, Moiseenko A, Shaitan K, Polshakov V, Nedbal L, Gorelova O. 2017. Versatility of the green microalga cell vacuole function as revealed by analytical transmission electron microscopy. Protoplasma 254:1323–1340. doi: 10.1007/s00709-016-1024-5 [DOI] [PubMed] [Google Scholar]
- 63. Yagisawa F, Nishida K, Yoshida M, Ohnuma M, Shimada T, Fujiwara T, Yoshida Y, Misumi O, Kuroiwa H, Kuroiwa T. 2009. Identification of novel proteins in isolated polyphosphate vacuoles in the primitive red alga Cyanidioschyzon merolae. Plant J 60:882–893. doi: 10.1111/j.1365-313X.2009.04008.x [DOI] [PubMed] [Google Scholar]
- 64. Schlatterer C, Buravkov S, Zierold K, Knoll G. 1994. Calcium-sequestering organelles of Dictyostelium discoideum: changes in element content during early development as measured by electron probe X-ray microanalysis. Cell Calcium 16:101–111. doi: 10.1016/0143-4160(94)90005-1 [DOI] [PubMed] [Google Scholar]
- 65. Schlatterer C, Walther P, Müller M, Mendgen K, Zierold K, Knoll G. 2001. Calcium stores in differentiated Dictyostelium discoideum: prespore cells sequester calcium more efficiently than prestalk cells. Cell Calcium 29:171–182. doi: 10.1054/ceca.2000.0181 [DOI] [PubMed] [Google Scholar]
- 66. MacDonald JIS, Weeks G. 1988. Evidence for a membrane-bound pyrophosphatase in Dictyostelium discoideum. FEBS Lett 238:9–12. doi: 10.1016/0014-5793(88)80214-x [DOI] [PubMed] [Google Scholar]
- 67. León G, Fiori C, Das P, Moreno M, Tovar R, Sánchez-Salas JL, Muñoz ML. 1997. Electron probe analysis and biochemical characterization of electron-dense granules secreted by Entamoeba histolytica. Mol Biochem Parasitol 85:233–242. doi: 10.1016/s0166-6851(97)02833-8 [DOI] [PubMed] [Google Scholar]
- 68. Gerasimaitė R, Sharma S, Desfougères Y, Schmidt A, Mayer A. 2014. Coupled synthesis and translocation restrains polyphosphate to acidocalcisome-like vacuoles and prevents its toxicity. J Cell Sci 127:5093–5104. doi: 10.1242/jcs.159772 [DOI] [PubMed] [Google Scholar]
- 69. Hovnanyan K, Marutyan S, Marutyan S, Hovnanyan M, Navasardyan L, Trchounian A. 2020. Ultrastructural investigation of acidocalcisomes and ATPase activity in yeast Candida guilliermondii NP-4 as 'complementary' stress-targets. Lett Appl Microbiol 71:413–419. doi: 10.1111/lam.13348 [DOI] [PubMed] [Google Scholar]
- 70. Milani G, Schereiber AZ, Vercesi AE. 2001. Ca2+ Transport into an intracellular acidic compartment of Candida parapsilosis. FEBS Lett 500:80–84. doi: 10.1016/s0014-5793(01)02585-6 [DOI] [PubMed] [Google Scholar]
- 71. Bowman BJ, Draskovic M, Freitag M, Bowman EJ. 2009. Structure and distribution of organelles and cellular location of calcium transporters in Neurospora crassa. Eukaryot Cell 8:1845–1855. doi: 10.1128/EC.00174-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Kikuchi Y, Hijikata N, Yokoyama K, Ohtomo R, Handa Y, Kawaguchi M, Saito K, Ezawa T. 2014. Polyphosphate accumulation is driven by transcriptome alterations that lead to near-synchronous and near-equivalent uptake of inorganic cations in an arbuscular mycorrhizal fungus. New Phytol 204:638–649. doi: 10.1111/nph.12937 [DOI] [PubMed] [Google Scholar]
- 73. Nguyen CT, Saito K. 2021. Role of cell wall polyphosphates in phosphorus transfer at the arbuscular interface in Mycorrhizas. Front Plant Sci 12:725939. doi: 10.3389/fpls.2021.725939 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74. Ramos IB, Miranda K, Ulrich P, Ingram P, LeFurgey A, Machado EA, de Souza W, Docampo R. 2010. Calcium- and polyphosphate-containing acidocalcisomes in chicken egg yolk. Biol Cell 102:421–434. doi: 10.1042/BC20100011 [DOI] [PubMed] [Google Scholar]
- 75. Miranda K, Benchimol M, Docampo R, de Souza W. 2000. The fine structure of acidocalcisomes in Trypanosoma cruzi. Parasitol Res 86:373–384. doi: 10.1007/s004360050682 [DOI] [PubMed] [Google Scholar]
- 76. Miranda K, Docampo R, Grillo O, Franzen A, Attias M, Vercesi A, Plattner H, Hentschel J, de Souza W. 2004. Dynamics of polymorphism of acidocalcisomes in Leishmania parasites. Histochem Cell Biol 121:407–418. doi: 10.1007/s00418-004-0646-4 [DOI] [PubMed] [Google Scholar]
- 77. Hughes L, Borrett S, Towers K, Starborg T, Vaughan S. 2017. Patterns of organelle ontogeny through a cell cycle revealed by whole-cell reconstructions using 3D electron microscopy. J Cell Sci 130:637–647. doi: 10.1242/jcs.198887 [DOI] [PubMed] [Google Scholar]
- 78. Girard-Dias W, Augusto I, V A Fernandes T, G Pascutti P, de Souza W, Miranda K. 2023. A spatially resolved elemental nanodomain organization within acidocalcisomes in Trypanosoma cruzi. Proc Natl Acad Sci U S A 120:e2300942120. doi: 10.1073/pnas.2300942120 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Moreno B, Bailey BN, Luo S, Martin MB, Kuhlenschmidt M, Moreno SN, Docampo R, Oldfield E. 2001. 31P NMR of apicomplexans and the effects of risedronate on Cryptosporidium parvum growth. Biochem Biophys Res Commun 284:632–637. doi: 10.1006/bbrc.2001.5009 [DOI] [PubMed] [Google Scholar]
- 80. Negreiros RS, Lander N, Huang G, Cordeiro CD, Smith SA, Morrissey JH, Docampo R. 2018. Inorganic polyphosphate interacts with nucleolar and glycosomal proteins in trypanosomatids. Mol Microbiol 110:973–994. doi: 10.1111/mmi.14131 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Rohloff P, Rodrigues CO, Docampo R. 2003. Regulatory volume decrease in Trypanosoma cruzi involves amino acid efflux and changes in intracellular calcium. Mol Biochem Parasitol 126:219–230. doi: 10.1016/s0166-6851(02)00277-3 [DOI] [PubMed] [Google Scholar]
- 82. Kakinuma Y, Masuda N, Igarashi K. 1992. Proton potential-dependent polyamine transport by vacuolar membrane vesicles of Saccharomyces cerevisiae . Biochim Biophys Acta 1107:126–130. doi: 10.1016/0005-2736(92)90337-l [DOI] [PubMed] [Google Scholar]
- 83. Tomitori H, Kashiwagi K, Asakawa T, Kakinuma Y, Michael AJ, Igarashi K. 2001. Multiple polyamine transport systems on the vacuolar membrane in yeast. Biochem J 353:681–688. doi: 10.1042/0264-6021:3530681 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Merchant SS, Schmollinger S, Strenkert D, Moseley JL, Blaby-Haas CE. 2020. From economy to luxury: copper homeostasis in Chlamydomonas and other algae. Biochim Biophys Acta Mol Cell Res 1867:118822. doi: 10.1016/j.bbamcr.2020.118822 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Tsednee M, Castruita M, Salomé PA, Sharma A, Lewis BE, Schmollinger SR, Strenkert D, Holbrook K, Otegui MS, Khatua K, Das S, Datta A, Chen S, Ramon C, Ralle M, Weber PK, Stemmler TL, Pett-Ridge J, Hoffman BM, Merchant SS. 2019. Manganese co-localizes with calcium and phosphorus in Chlamydomonas acidocalcisomes and is mobilized in manganese-deficient conditions. J Biol Chem 294:17626–17641. doi: 10.1074/jbc.RA119.009130 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Long H, Fang J, Ye L, Zhang B, Hui C, Deng X, Merchant SS, Huang K. 2023. Structural and functional regulation of Chlamydomonas lysosome-related organelles during environmental changes. Plant Physiol 192:927–944. doi: 10.1093/plphys/kiad189 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Nguyen TQ, Dziuba N, Lindahl PA. 2019. Isolated Saccharomyces cerevisiae vacuoles contain low-molecular-mass transition-metal polyphosphate complexes. Metallomics 11:1298–1309. doi: 10.1039/c9mt00104b [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88. Salto ML, Kuhlenschmidt T, Kuhlenschmidt M, de Lederkremer RM, Docampo R. 2008. Phospholipid and glycolipid composition of acidocalcisomes of Trypanosoma cruzi. Mol Biochem Parasitol 158:120–130. doi: 10.1016/j.molbiopara.2007.12.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89. Billington K, Halliday C, Madden R, Dyer P, Barker AR, Moreira-Leite FF, Carrington M, Vaughan S, Hertz-Fowler C, Dean S, Sunter JD, Wheeler RJ, Gull K. 2023. Genome-wide subcellular protein map for the flagellate parasite Trypanosoma brucei. Nat Microbiol 8:533–547. doi: 10.1038/s41564-022-01295-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Sunter JD, Dean S, Wheeler RJ. 2023. TrypTag.org: from images to discoveries using genome-wide protein localisation in Trypanosoma brucei. Trends in Parasitology 39:328–331. doi: 10.1016/j.pt.2023.02.008 [DOI] [PubMed] [Google Scholar]
- 91. Fang J, Rohloff P, Miranda K, Docampo R. 2007. Ablation of a small transmembrane protein of Trypanosoma brucei (TbVTC1) involved in the synthesis of polyphosphate alters acidocalcisome biogenesis and function, and leads to a cytokinesis defect. Biochem J 407:161–170. doi: 10.1042/BJ20070612 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Lander N, Ulrich PN, Docampo R. 2013. Trypanosoma brucei vacuolar transporter chaperone 4 (TbVtc4) is an acidocalcisome polyphosphate kinase required for in vivo infection. J Biol Chem 288:34205–34216. doi: 10.1074/jbc.M113.518993 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Ulrich PN, Lander N, Kurup SP, Reiss L, Brewer J, Soares Medeiros LC, Miranda K, Docampo R. 2014. The acidocalcisome vacuolar transporter chaperone 4 catalyzes the synthesis of polyphosphate in insect-stages of Trypanosoma brucei and T. cruzi. J Eukaryot Microbiol 61:155–165. doi: 10.1111/jeu.12093 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. Isah MB, Goldring JPD, Coetzer THT. 2020. Expression and copper binding properties of the N-terminal domain of copper P-type ATPases of African trypanosomes. Mol Biochem Parasitol 235:111245. doi: 10.1016/j.molbiopara.2019.111245 [DOI] [PubMed] [Google Scholar]
- 95. Paul R, Banerjee S, Sen S, Dubey P, Maji S, Bachhawat AK, Datta R, Gupta A. 2022. A novel leishmanial copper P-type ATPase plays a vital role in parasite infection and intracellular survival. J Biol Chem 298:101539. doi: 10.1016/j.jbc.2021.101539 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96. Luo S, Rohloff P, Cox J, Uyemura SA, Docampo R. 2004. Trypanosoma brucei plasma membrane-type Ca2+-ATPase 1 (TbPMC1) and 2 (TbPMC2) genes encode functional Ca2+-ATPases localized to the acidocalcisomes and plasma membrane, and essential for Ca2+ homeostasis and growth. J Biol Chem 279:14427–14439. doi: 10.1074/jbc.M309978200 [DOI] [PubMed] [Google Scholar]
- 97. Huang G, Bartlett PJ, Thomas AP, Moreno SNJ, Docampo R. 2013. Acidocalcisomes of Trypanosoma brucei have an inositol 1,4,5-trisphosphate receptor that is required for growth and infectivity. Proc Natl Acad Sci U S A 110:1887–1892. doi: 10.1073/pnas.1216955110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Steinmann ME, Schmidt RS, Bütikofer P, Mäser P, Sigel E. 2017. TbIRK is a signature sequence free potassium channel from Trypanosoma brucei locating to acidocalcisomes. Sci Rep 7:656. doi: 10.1038/s41598-017-00752-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99. Lemercier G, Espiau B, Ruiz FA, Vieira M, Luo S, Baltz T, Docampo R, Bakalara N. 2004. A pyrophosphatase regulating polyphosphate metabolism in acidocalcisomes is essential for Trypanosoma brucei virulence in mice. J Biol Chem 279:3420–3425. doi: 10.1074/jbc.M309974200 [DOI] [PubMed] [Google Scholar]
- 100. Huang G, Fang J, Sant’Anna C, Li Z-H, Wellems DL, Rohloff P, Docampo R. 2011. Adaptor protein-3 (AP-3) complex mediates the biogenesis of acidocalcisomes and is essential for growth and virulence of Trypanosoma brucei. J Biol Chem 286:36619–36630. doi: 10.1074/jbc.M111.284661 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Emmer BT, Nakayasu ES, Souther C, Choi H, Sobreira TJP, Epting CL, Nesvizhskii AI, Almeida IC, Engman DM. 2011. Global analysis of protein palmitoylation in African trypanosomes. Eukaryot Cell 10:455–463. doi: 10.1128/EC.00248-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Sun SY, Wang C, Yuan YA, He CY. 2013. An intracellular membrane junction consisting of flagellum adhesion glycoproteins links flagellum biogenesis to cell morphogenesis in Trypanosoma brucei. J Cell Sci 126:520–531. doi: 10.1242/jcs.113621 [DOI] [PubMed] [Google Scholar]
- 103. Ferella M, Nilsson D, Darban H, Rodrigues C, Bontempi EJ, Docampo R, Andersson B. 2008. Proteomics in Trypanosoma cruzi--localization of novel proteins to various organelles. Proteomics 8:2735–2749. doi: 10.1002/pmic.200700940 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Lu H-G, Zhong L, de Souza W, Benchimol M, Moreno S, Docampo R. 1998. Ca2+ content and expression of an acidocalcisomal calcium pump are elevated in intracellular forms of Trypanosoma cruzi . Mol Cell Biol 18:2309–2323. doi: 10.1128/MCB.18.4.2309 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105. Montalvetti A, Rohloff P, Docampo R. 2004. A functional aquaporin co-localizes with the vacuolar proton pyrophosphatase to acidocalcisomes and the contractile vacuole complex of Trypanosoma cruzi. J Biol Chem 279:38673–38682. doi: 10.1074/jbc.M406304200 [DOI] [PubMed] [Google Scholar]
- 106. Rohloff P, Montalvetti A, Docampo R. 2004. Acidocalcisomes and the contractile vacuole complex are involved in osmoregulation in Trypanosoma cruzi . J Biol Chem 279:52270–52281. doi: 10.1074/jbc.M410372200 [DOI] [PubMed] [Google Scholar]
- 107. Lander N, Chiurillo MA, Storey M, Vercesi AE, Docampo R. 2016. CRISPR/Cas9-mediated endogenous C-terminal tagging of Trypanosoma cruzi genes reveals the acidocalcisome localization of the inositol 1,4,5-trisphosphate receptor. J Biol Chem 291:25505–25515. doi: 10.1074/jbc.M116.749655 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Chiurillo MA, Lander N, Vercesi AE, Docampo R. 2020. Ip3 receptor-mediated Ca2+ release from acidocalcisomes regulates mitochondrial bioenergetics and prevents autophagy in Trypanosoma cruzi. Cell Calcium 92:102284. doi: 10.1016/j.ceca.2020.102284 [DOI] [PubMed] [Google Scholar]
- 109. Galizzi M, Bustamante JM, Fang J, Miranda K, Soares Medeiros LC, Tarleton RL, Docampo R. 2013. Evidence for the role of vacuolar soluble pyrophosphatase and inorganic polyphosphate in Trypanosoma cruzi persistence. Mol Microbiol 90:699–715. doi: 10.1111/mmi.12392 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110. Batista CM, Saad F, Ceccoti SPC, Eger I, Soares MJ. 2018. Subcellular localisation of FLAG tagged enzymes of the dynamic protein S-palmitoylation cycle of Trypanosoma cruzi epimastigotes. Mem Inst Oswaldo Cruz 113:e180086. doi: 10.1590/0074-02760180086 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111. Mantilla BS, Azevedo C, Denny PW, Saiardi A, Docampo R. 2021. The histidine ammonia lyase of Trypanosoma cruzi is involved in acidocalcisome alkalinization and is essential for survival under starvation conditions. mBio 12:e01981-21. doi: 10.1128/mBio.01981-21 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112. Luo S, Marchesini N, Moreno SN, Docampo R. 1999. A plant-like vacuolar H+-pyrophosphatase in Plasmodium falciparum. FEBS Lett 460:217–220. doi: 10.1016/s0014-5793(99)01353-8 [DOI] [PubMed] [Google Scholar]
- 113. Chasen NM, Stasic AJ, Asady B, Coppens I, Moreno SNJ. 2019. The vacuolar zinc transporter TgZnT protects Toxoplasma gondii from zinc toxicity. mSphere 4:e00086-19. doi: 10.1128/mSphere.00086-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. Kenthirapalan S, Waters AP, Matuschewski K, Kooij TWA. 2014. Copper-transporting ATPase is important for malaria parasite fertility. Mol Microbiol 91:315–325. doi: 10.1111/mmi.12461 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115. Liu J, Pace D, Dou Z, King TP, Guidot D, Li Z-H, Carruthers VB, Moreno SNJ. 2014. A vacuolar-H(+) -pyrophosphatase (TgVP1) is required for microneme secretion, host cell invasion, and extracellular survival of Toxoplasma gondii. Mol Microbiol 93:698–712. doi: 10.1111/mmi.12685 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Stasic AJ, Chasen NM, Dykes EJ, Vella SA, Asady B, Starai VJ, Moreno SNJ. 2019. The toxoplasma vacuolar H+-ATPase regulates intracellular pH and impacts the maturation of essential secretory proteins. Cell Rep 27:2132–2146. doi: 10.1016/j.celrep.2019.04.038 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117. Benchimol M, De Souza W, Vanderheyden N, Zhong L, Lu HG, Moreno SN, Docampo R. 1998. Functional expression of a vacuolar-type H+-ATPase in the plasma membrane and intracellular vacuoles of Trypanosoma cruzi. Biochem J 332 (Pt 3):695–702. doi: 10.1042/bj3320695 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118. Stasic AJ, Dykes EJ, Cordeiro CD, Vella SA, Fazli MS, Quinn S, Docampo R, Moreno SNJ. 2021. Ca2+ entry at the plasma membrane and uptake by acidic stores is regulated by the activity of the V-H+ -ATPase in Toxoplasma gondii. Mol Microbiol 115:1054–1068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Hayashi M, Yamada H, Mitamura T, Horii T, Yamamoto A, Moriyama Y. 2000. Vacuolar H+-ATPase localized in plasma membranes of malaria parasite cells, Plasmodium falciparum, is involved in regional acidification of parasitized erythrocytes. J Biol Chem 275:34353–34358. doi: 10.1074/jbc.M003323200 [DOI] [PubMed] [Google Scholar]
- 120. Saliba KJ, Allen RJW, Zissis S, Bray PG, Ward SA, Kirk K. 2003. Acidification of the malaria parasite’s digestive vacuole by a H+-ATPase and a H+-pyrophosphatase. J Biol Chem 278:5605–5612. doi: 10.1074/jbc.M208648200 [DOI] [PubMed] [Google Scholar]
- 121. Elandalloussi LM, Adams B, Smith PJ. 2005. ATPase activity of purified plasma membranes and digestive vacuoles from Plasmodium falciparum. Mol Biochem Parasitol 141:49–56. doi: 10.1016/j.molbiopara.2005.02.001 [DOI] [PubMed] [Google Scholar]
- 122. Saliba KJ, Kirk K. 1999. pH regulation in the intracellular malaria parasite, Plasmodium falciparum. H+ extrusion via a V-type H+-ATPase. J Biol Chem 274:33213–33219. doi: 10.1074/jbc.274.47.33213 [DOI] [PubMed] [Google Scholar]
- 123. Rea PA, Kim Y, Sarafian V, Poole RJ, Davies JM, Sanders D. 1992. Vacuolar H+-translocating pyrophosphatases: a new category of ion translocase. Trends Biochem Sci 17:348–353. doi: 10.1016/0968-0004(92)90313-x [DOI] [PubMed] [Google Scholar]
- 124. Lin SM, Tsai JY, Hsiao CD, Huang YT, Chiu CL, Liu MH, Tung JY, Liu TH, Pan RL, Sun YJ. 2012. Crystal structure of a membrane-embedded H+-translocating pyrophosphatase. Nature 484:399–403. doi: 10.1038/nature10963 [DOI] [PubMed] [Google Scholar]
- 125. Martinez R, Wang Y, Benaim G, Benchimol M, de Souza W, Scott DA, Docampo R. 2002. A proton pumping pyrophosphatase in the Golgi apparatus and plasma membrane vesicles of Trypanosoma cruzi. Mol Biochem Parasitol 120:205–213. doi: 10.1016/s0166-6851(01)00456-x [DOI] [PubMed] [Google Scholar]
- 126. Hill JE, Scott DA, Luo S, Docampo R. 2000. Cloning and functional expression of a gene encoding a vacuolar-type proton-translocating pyrophosphatase from Trypanosoma cruzi. Biochem J 351:281–288. doi: 10.1042/0264-6021:3510281 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127. Lemercier G, Dutoya S, Luo S, Ruiz FA, Rodrigues CO, Baltz T, Docampo R, Bakalara N. 2002. A vacuolar-type H+-pyrophosphatase governs maintenance of functional acidocalcisomes and growth of the insect and mammalian forms of Trypanosoma brucei. J Biol Chem 277:37369–37376. doi: 10.1074/jbc.M204744200 [DOI] [PubMed] [Google Scholar]
- 128. McIntosh MT, Drozdowicz YM, Laroiya K, Rea PA, Vaidya AB. 2001. Two classes of plant-like vacuolar-type H+-pyrophosphatases in malaria parasites. Mol Biochem Parasitol 114:183–195. doi: 10.1016/s0166-6851(01)00251-1 [DOI] [PubMed] [Google Scholar]
- 129. Miranda K, Pace DA, Cintron R, Rodrigues JCF, Fang J, Smith A, Rohloff P, Coelho E, de Haas F, de Souza W, Coppens I, Sibley LD, Moreno SNJ. 2010. Characterization of a novel organelle in Toxoplasma gondii with similar composition and function to the plant vacuole. Mol Microbiol 76:1358–1375. doi: 10.1111/j.1365-2958.2010.07165.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130. Segami S, Asaoka M, Kinoshita S, Fukuda M, Nakanishi Y, Maeshima M. 2018. Biochemical, structural and physiological characteristics of vacuolar H+-pyrophosphatase. Plant Cell Physiol 59:1300–1308. doi: 10.1093/pcp/pcy054 [DOI] [PubMed] [Google Scholar]
- 131. Potapenko E, Cordeiro CD, Huang G, Storey M, Wittwer C, Dutta AK, Jessen HJ, Starai VJ, Docampo R. 2018. 5-Diphosphoinositol pentakisphosphate (5-IP7) regulates phosphate release from acidocalcisomes and yeast vacuoles. J Biol Chem 293:19101–19112. doi: 10.1074/jbc.RA118.005884 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132. Kornberg A. 1962. On the metabolic significance of phosphorolytic and pyrophosphorolytic reactions, p 251–254. In Kasha M, Pullman B (ed), Horizons in Biochemistry. Academic Press, New York. [Google Scholar]
- 133. Rocha Facanha A, de Meis L. 1998. Reversibility of H+-ATPase and H+-pyrophosphatase in tonoplast vesicles from maize coleoptiles and seeds. Plant Physiol 116:1487–1495. doi: 10.1104/pp.116.4.1487 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134. Hirata T, Nakamura N, Omote H, Wada Y, Futai M. 2000. Regulation and reversibility of vacuolar H+-ATPase. J Biol Chem 275:386–389. doi: 10.1074/jbc.275.1.386 [DOI] [PubMed] [Google Scholar]
- 135. Moniakis J, Coukell MB, Forer A. 1995. Molecular cloning of an intracellular P-type ATPase from Dictyostelium that is up-regulated in calcium-adapted cells. J Biol Chem 270:28276–28281. doi: 10.1074/jbc.270.47.28276 [DOI] [PubMed] [Google Scholar]
- 136. Cunningham KW, Fink GR. 1994. Calcineurin-dependent growth control in Saccharomyces cerevisiae mutants lacking PMC1, a homolog of plasma membrane Ca2+ ATPases. J Cell Biol 124:351–363. doi: 10.1083/jcb.124.3.351 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137. Ghosh SK, Rosenthal B, Rogers R, Samuelson J. 2000. Vacuolar localization of an Entamoeba histolytica homologue of the plasma membrane ATPase (PMCA). Mol Biochem Parasitol 108:125–130. doi: 10.1016/s0166-6851(00)00196-1 [DOI] [PubMed] [Google Scholar]
- 138. Pérez-Gordones MC, Ramírez-Iglesias JR, Cervino V, Uzcanga GL, Benaim G, Mendoza M. 2017. Evidence of the presence of a calmodulin-sensitive plasma membrane Ca2+-ATPase in Trypanosoma equiperdum. Mol Biochem Parasitol 213:1–11. doi: 10.1016/j.molbiopara.2017.02.001 [DOI] [PubMed] [Google Scholar]
- 139. Benaim G, Losada S, Gadelha FR, Docampo R. 1991. A calmodulin-activated (Ca2+-Mg2+)-ATPase is involved in Ca2+ transport by plasma membrane vesicles from Trypanosoma cruzi. Biochem J 280 (Pt 3):715–720. doi: 10.1042/bj2800715 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140. Hasan NM, Lutsenko S. 2012. Regulation of copper transporters in human cells. Curr Top Membr 69:137–161. doi: 10.1016/B978-0-12-394390-3.00006-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141. Hong-Hermesdorf A, Miethke M, Gallaher SD, Kropat J, Dodani SC, Chan J, Barupala D, Domaille DW, Shirasaki DI, Loo JA, Weber PK, Pett-Ridge J, Stemmler TL, Chang CJ, Merchant SS. 2014. Subcellular metal imaging identifies dynamic sites of Cu accumulation in Chlamydomonas. Nat Chem Biol 10:1034–1042. doi: 10.1038/nchembio.1662 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142. Hothorn M, Neumann H, Lenherr ED, Wehner M, Rybin V, Hassa PO, Uttenweiler A, Reinhardt M, Schmidt A, Seiler J, Ladurner AG, Herrmann C, Scheffzek K, Mayer A. 2009. Catalytic core of a membrane-associated eukaryotic polyphosphate polymerase. Science 324:513–516. doi: 10.1126/science.1168120 [DOI] [PubMed] [Google Scholar]
- 143. Wild R, Gerasimaite R, Jung JY, Truffault V, Pavlovic I, Schmidt A, Saiardi A, Jessen HJ, Poirier Y, Hothorn M, Mayer A. 2016. Control of eukaryotic phosphate homeostasis by Inositol polyphosphate sensor domains. Science 352:986–990. doi: 10.1126/science.aad9858 [DOI] [PubMed] [Google Scholar]
- 144. Guan Z, Chen J, Liu R, Chen Y, Xing Q, Du Z, Cheng M, Hu J, Zhang W, Mei W, Wan B, Wang Q, Zhang J, Cheng P, Cai H, Cao J, Zhang D, Yan J, Yin P, Hothorn M, Liu Z. 2023. The cytoplasmic synthesis and coupled membrane translocation of eukaryotic polyphosphate by signal-activated VTC complex. Nat Commun 14:718. doi: 10.1038/s41467-023-36466-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145. Liu W, Wang J, Comte-Miserez V, Zhang M, Yu X, Chen Q, Jessen HJ, Mayer A, Wu S, Ye S. 2023. Cryo-EM structure of the polyphosphate polymerase VTC reveals coupling of polymer synthesis to membrane transit. EMBO J 42:e113320. doi: 10.15252/embj.2022113320 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146. Pipercevic J, Kohl B, Gerasimaite R, Comte-Miserez V, Hostachy S, Müntener T, Agustoni E, Jessen HJ, Fiedler D, Mayer A, Hiller S. 2023. Inositol pyrophosphates activate the vacuolar transport chaperone complex in yeast by disrupting a homotypic SPX domain interaction. Nat Commun 14:2645. doi: 10.1038/s41467-023-38315-w [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147. Klompmaker SH, Kohl K, Fasel N, Mayer A. 2017. Magnesium uptake by connecting fluid-phase endocytosis to an intracellular inorganic cation filter. Nat Commun 8:1879. doi: 10.1038/s41467-017-01930-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148. Kohl K, Zangger H, Rossi M, Isorce N, Lye LF, Owens KL, Beverley SM, Mayer A, Fasel N. 2018. Importance of polyphosphate in the Leishmania life cycle. Microb Cell 5:371–384. doi: 10.15698/mic2018.08.642 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149. Murray JM, Johnson DI. 2001. The Cdc42p GTPase and its regulators Nrf1p and Scd1p are involved in endocytic trafficking in the fission yeast Schizosaccharomyces pombe . J Biol Chem 276:3004–3009. doi: 10.1074/jbc.M007389200 [DOI] [PubMed] [Google Scholar]
- 150. Johnson DI. 1999. Cdc42: an essential Rho-type GTPase controlling eukaryotic cell polarity. Microbiol Mol Biol Rev 63:54–105. doi: 10.1128/MMBR.63.1.54-105.1999 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151. Richman TJ, Sawyer MM, Johnson DI. 2002. Saccharomyces cerevisiae Cdc42p localizes to cellular membranes and clusters at sites of polarized growth. Eukaryot Cell 1:458–468. doi: 10.1128/EC.1.3.458-468.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152. Müller O, Johnson DI, Mayer A. 2001. Cdc42p functions at the docking stage of yeast vacuole membrane fusion. EMBO J 20:5657–5665. doi: 10.1093/emboj/20.20.5657 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153. Rooney PJ, Ayong L, Tobin CM, Moreno SNJ, Knoll LJ. 2011. TgVTC2 is involved in polyphosphate accumulation in Toxoplasma gondii. Mol Biochem Parasitol 176:121–126. doi: 10.1016/j.molbiopara.2010.12.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154. Aksoy M, Pootakham W, Grossman AR. 2014. Critical function of a Chlamydomonas reinhardtii putative polyphosphate polymerase subunit during nutrient deprivation. Plant Cell 26:4214–4229. doi: 10.1105/tpc.114.129270 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155. Wang L, Jia X, Zhang Y, Xu L, Menand B, Zhao H, Zeng H, Dolan L, Zhu Y, Yi K. 2021. Loss of two families of SPX domain-containing proteins required for vacuolar polyphosphate accumulation coincides with the transition to phosphate storage in green plants. Mol Plant 14:838–846. doi: 10.1016/j.molp.2021.01.015 [DOI] [PubMed] [Google Scholar]
- 156. Plouviez M, Fernández E, Grossman AR, Sanz-Luque E, Sells M, Wheeler D, Guieysse B. 2021. Responses of Chlamydomonas reinhardtii during the transition from P-deficient to P-sufficient growth (the P-overplus response): the roles of the vacuolar transport chaperones and polyphosphate synthesis. J Phycol 57:988–1003. doi: 10.1111/jpy.13145 [DOI] [PubMed] [Google Scholar]
- 157. Kumble KD, Kornberg A. 1996. Endopolyphosphatases for long chain inorganic polyphosphate in yeast and mammals. J Biol Chem 271:27146–27151. doi: 10.1074/jbc.271.43.27146 [DOI] [PubMed] [Google Scholar]
- 158. Shi X, Kornberg A. 2005. Endopolyphosphatase in Saccharomyces cerevisiae undergoes post-translational activations to produce short-chain polyphosphates. FEBS Lett 579:2014–2018. doi: 10.1016/j.febslet.2005.02.032 [DOI] [PubMed] [Google Scholar]
- 159. Sethuraman A, Rao NN, Kornberg A. 2001. The endopolyphosphatase gene: essential in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 98:8542–8547. doi: 10.1073/pnas.151269398 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160. Andreeva N, Trilisenko L, Eldarov M, Kulakovskaya T. 2015. Polyphosphatase PPN1 of Saccharomyces cerevisiae: switching of exopolyphosphatase and endopolyphosphatase activities. PLoS One 10:e0119594. doi: 10.1371/journal.pone.0119594 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161. Gerasimaitė R, Mayer A. 2017. Ppn2, a novel Zn2+-dependent polyphosphatase in the acidocalcisome-like yeast vacuole. J Cell Sci 130:1625–1636. doi: 10.1242/jcs.201061 [DOI] [PubMed] [Google Scholar]
- 162. Espiau B, Lemercier G, Ambit A, Bringaud F, Merlin G, Baltz T, Bakalara N. 2006. A soluble pyrophosphatase, a key enzyme for polyphosphate metabolism in Leishmania . J Biol Chem 281:1516–1523. doi: 10.1074/jbc.M506947200 [DOI] [PubMed] [Google Scholar]
- 163. Kotsikorou E, Song Y, Chan JMW, Faelens S, Tovian Z, Broderick E, Bakalara N, Docampo R, Oldfield E. 2005. Bisphosphonate inhibition of the exopolyphosphatase activity of the Trypanosoma brucei soluble vacuolar pyrophosphatase. J Med Chem 48:6128–6139. doi: 10.1021/jm058220g [DOI] [PubMed] [Google Scholar]
- 164. Yang Y, Ko T-P, Chen C-C, Huang G, Zheng Y, Liu W, Wang I, Ho M-R, Hsu S-TD, O’Dowd B, Huff HC, Huang C-H, Docampo R, Oldfield E, Guo R-T. 2016. Structures of trypanosome vacuolar soluble pyrophosphatases: antiparasitic drug targets. ACS Chem Biol 11:1362–1371. doi: 10.1021/acschembio.5b00724 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165. Jamwal A, Round AR, Bannwarth L, Venien-Bryan C, Belrhali H, Yogavel M, Sharma A. 2015. Structural and functional highlights of vacuolar soluble protein 1 from pathogen Trypanosoma brucei brucei. J Biol Chem 290:30498–30513. doi: 10.1074/jbc.M115.674176 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166. Li Z-H, Alvarez VE, De Gaudenzi JG, Sant’Anna C, Frasch ACC, Cazzulo JJ, Docampo R. 2011. Hyperosmotic stress induces aquaporin-dependent cell shrinkage, polyphosphate synthesis, amino acid accumulation, and global gene expression changes in Trypanosoma cruzi. J Biol Chem 286:43959–43971. doi: 10.1074/jbc.M111.311530 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167. Niyogi S, Jimenez V, Girard-Dias W, de Souza W, Miranda K, Docampo R. 2015. Rab32 is essential for maintaining functional acidocalcisomes, and for growth and infectivity of Trypanosoma cruzi. J Cell Sci 128:2363–2373. doi: 10.1242/jcs.169466 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168. Ulrich PN, Jimenez V, Park M, Martins VP, Atwood J, Moles K, Collins D, Rohloff P, Tarleton R, Moreno SNJ, Orlando R, Docampo R, Langsley G. 2011. Identification of contractile vacuole proteins in Trypanosoma cruzi. PLoS ONE 6:e18013. doi: 10.1371/journal.pone.0018013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169. Prole DL, Taylor CW. 2011. Identification of intracellular and plasma membrane calcium channel homologues in pathogenic parasites. PLoS One 6:e26218. doi: 10.1371/journal.pone.0026218 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170. Li F-J, Tan KSW, He CY. 2017. BAPTA-AM decreases cellular pH, inhibits acidocalcisome acidification and autophagy in amino acid-starved T. brucei. Mol Biochem Parasitol 213:26–29. doi: 10.1016/j.molbiopara.2017.03.001 [DOI] [PubMed] [Google Scholar]
- 171. Hashimoto M, Enomoto M, Morales J, Kurebayashi N, Sakurai T, Hashimoto T, Nara T, Mikoshiba K. 2013. Inositol 1,4,5-trisphosphate receptor regulates replication, differentiation, infectivity and virulence of the parasitic protist Trypanosoma cruzi. Mol Microbiol 87:1133–1150. doi: 10.1111/mmi.12155 [DOI] [PubMed] [Google Scholar]
- 172. Potapenko E, Negrão NW, Huang G, Docampo R. 2019. The acidocalcisome inositol-1,4,5-trisphosphate receptor of Trypanosoma brucei is stimulated by luminal polyphosphate hydrolysis products. J Biol Chem 294:10628–10637. doi: 10.1074/jbc.RA119.007906 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173. Palmer CP, Zhou XL, Lin J, Loukin SH, Kung C, Saimi Y. 2001. A TRP homolog in Saccharomyces cerevisiae forms an intracellular Ca2+-permeable channel in the yeast vacuolar membrane. Proc Natl Acad Sci U S A 98:7801–7805. doi: 10.1073/pnas.141036198 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174. Denis V, Cyert MS. 2002. Internal Ca2+ release in yeast is triggered by hypertonic shock and mediated by a TRP channel homologue. J Cell Biol 156:29–34. doi: 10.1083/jcb.200111004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175. Ahmed T, Nisler CR, Fluck EC, Walujkar S, Sotomayor M, Moiseenkova-Bell VY. 2022. Structure of the ancient TRPY1 channel from Saccharomyces cerevisiae reveals mechanisms of modulation by lipids and calcium. Structure 30:139–155. doi: 10.1016/j.str.2021.08.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176. Nies DH, Silver S. 1995. Ion efflux systems involved in bacterial metal resistances. J Ind Microbiol 14:186–199. doi: 10.1007/BF01569902 [DOI] [PubMed] [Google Scholar]
- 177. Kawachi M, Kobae Y, Kogawa S, Mimura T, Krämer U, Maeshima M. 2012. Amino acid screening based on structural modeling identifies critical residues for the function, ion selectivity and structure of Arabidopsis MTP1. FEBS J 279:2339–2356. doi: 10.1111/j.1742-4658.2012.08613.x [DOI] [PubMed] [Google Scholar]
- 178. Maret W. 2017. Zinc in cellular regulation: the nature and significance of "zinc signals. Int J Mol Sci 18:2285. doi: 10.3390/ijms18112285 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179. Kim SA, Punshon T, Lanzirotti A, Li L, Alonso JM, Ecker JR, Kaplan J, Guerinot ML. 2006. Localization of iron in Arabidopsis seed requires the vacuolar membrane transporter VIT1. Science 314:1295–1298. doi: 10.1126/science.1132563 [DOI] [PubMed] [Google Scholar]
- 180. Li L, Chen OS, McVey Ward D, Kaplan J. 2001. CCC1 is a transporter that mediates vacuolar iron storage in yeast. J Biol Chem 276:29515–29519. doi: 10.1074/jbc.M103944200 [DOI] [PubMed] [Google Scholar]
- 181. Slavic K, Krishna S, Lahree A, Bouyer G, Hanson KK, Vera I, Pittman JK, Staines HM, Mota MM. 2016. A vacuolar iron-transporter homologue acts as a detoxifier in Plasmodium. Nat Commun 7:10403. doi: 10.1038/ncomms10403 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182. Hürlimann HC, Stadler-Waibel M, Werner TP, Freimoser FM, Subramani S. 2007. Pho91 is a vacuolar phosphate transporter that regulates phosphate and polyphosphate metabolism in Saccharomyces cerevisiae. Mol Biol Cell 18:4438–4445. doi: 10.1091/mbc.e07-05-0457 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183. Jimenez V, Docampo R. 2015. TcPho91 is a contractile vacuole phosphate sodium symporter that regulates phosphate and polyphosphate metabolism in Trypanosoma cruzi. Mol Microbiol 97:911–925. doi: 10.1111/mmi.13075 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184. Vercesi AE, Docampo R. 1996. Sodium-proton exchange stimulates Ca2+ release from acidocalcisomes of Trypanosoma brucei. Biochem J 315 (Pt 1):265–270. doi: 10.1042/bj3150265 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185. Vercesi EA, Grijalba TM, Docampo R. 1997. Inhibition of Ca2+ Release from Trypanosoma brucei acidocalcisomes by 3,5-dibutyl-4-hydroxytoluene: role of the Na+/H+ exchanger. Biochem J 328:479–482. doi: 10.1042/bj3280479 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186. Vercesi AE, Rodrigues CO, Catisti R, Docampo R. 2000. Presence of a Na+/H+ exchanger in acidocalcisomes of Leishmania donovani and their alkalization by anti-leishmanial drugs. FEBS Lett 473:203–206. doi: 10.1016/s0014-5793(00)01531-3 [DOI] [PubMed] [Google Scholar]
- 187. Cutler DF. 2002. Introduction: lysosome-related organelles. Semin Cell Dev Biol 13:261–262. doi: 10.1016/s108495210200054x [DOI] [PubMed] [Google Scholar]
- 188. Dell’Angelica EC, Mullins C, Caplan S, Bonifacino JS. 2000. Lysosome-related organelles. FASEB J 14:1265–1278. doi: 10.1096/fj.14.10.1265 [DOI] [PubMed] [Google Scholar]
- 189. Besteiro S, Tonn D, Tetley L, Coombs GH, Mottram JC. 2008. The AP3 adaptor is involved in the transport of membrane proteins to acidocalcisomes of Leishmania. J Cell Sci 121:561–570. doi: 10.1242/jcs.022574 [DOI] [PubMed] [Google Scholar]
- 190. Boehm M, Bonifacino JS. 2002. Genetic analyses of adaptin function from yeast to mammals. Gene 286:175–186. doi: 10.1016/s0378-1119(02)00422-5 [DOI] [PubMed] [Google Scholar]
- 191. Ihrke G, Kyttälä A, Russell MRG, Rous BA, Luzio JP. 2004. Differential use of two AP-3-mediated pathways by lysosomal membrane proteins. Traffic 5:946–962. doi: 10.1111/j.1600-0854.2004.00236.x [DOI] [PubMed] [Google Scholar]
- 192. Peden AA, Oorschot V, Hesser BA, Austin CD, Scheller RH, Klumperman J. 2004. Localization of the AP-3 adaptor complex defines a novel endosomal exit site for lysosomal membrane proteins. J Cell Biol 164:1065–1076. doi: 10.1083/jcb.200311064 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Bonifacino JS, Traub LM. 2003. Signals for sorting of transmembrane proteins to endosomes and lysosomes. Annu Rev Biochem 72:395–447. doi: 10.1146/annurev.biochem.72.121801.161800 [DOI] [PubMed] [Google Scholar]
- 194. Huizing M, Malicdan MCV, Wang JA, Pri-Chen H, Hess RA, Fischer R, O’Brien KJ, Merideth MA, Gahl WA, Gochuico BR. 2020. Hermansky-Pudlak syndrome: mutation update. Hum Mutat 41:543–580. doi: 10.1002/humu.23968 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195. Bentley-DeSousa A, Downey M. 2021. Vtc5 is localized to the vacuole membrane by the conserved AP-3 complex to regulate polyphosphate synthesis in budding yeast. mBio 12:e00994-21. doi: 10.1128/mBio.00994-21 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196. Kantheti P, Qiao X, Diaz ME, Peden AA, Meyer GE, Carskadon SL, Kapfhamer D, Sufalko D, Robinson MS, Noebels JL, Burmeister M. 1998. Mutation in AP-3 Delta in the mocha mouse links endosomal transport to storage deficiency in platelets, melanosomes, and synaptic vesicles. Neuron 21:111–122. doi: 10.1016/S0896-6273(00)80519-X [DOI] [PubMed] [Google Scholar]
- 197. Raposo G, Marks MS, Cutler DF. 2007. Lysosome-related organelles: driving post-Golgi compartments into specialisation. Curr Opin Cell Biol 19:394–401. doi: 10.1016/j.ceb.2007.05.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198. Zhang K, Hsu FF, Scott DA, Docampo R, Turk J, Beverley SM. 2005. Leishmania salvage and remodelling of host sphingolipids in amastigote survival and acidocalcisome biogenesis. Mol Microbiol 55:1566–1578. doi: 10.1111/j.1365-2958.2005.04493.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199. Madeira da Silva L, Beverley SM. 2010. Expansion of the target of rapamycin (TOR) kinase family and function in Leishmania shows that TOR3 is required for acidocalcisome biogenesis and animal infectivity. Proc Natl Acad Sci U S A 107:11965–11970. doi: 10.1073/pnas.1004599107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200. de Jesus TCL, Tonelli RR, Nardelli SC, da Silva Augusto L, Motta MCM, Girard-Dias W, Miranda K, Ulrich P, Jimenez V, Barquilla A, Navarro M, Docampo R, Schenkman S. 2010. Target of rapamycin (TOR)-like 1 kinase is involved in the control of polyphosphate levels and acidocalcisome maintenance in Trypanosoma brucei. J Biol Chem 285:24131–24140. doi: 10.1074/jbc.M110.120212 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201. Dutoya S, Gibert S, Lemercier G, Santarelli X, Baltz D, Baltz T, Bakalara N. 2001. A novel C-terminal Kinesin is essential for maintaining functional acidocalcisomes in Trypanosoma brucei. J Biol Chem 276:49117–49124. doi: 10.1074/jbc.M105962200 [DOI] [PubMed] [Google Scholar]
- 202. Ramakrishnan S, Asady B, Docampo R. 2018. Acidocalcisome-mitochondrion membrane contact sites in Trypanosoma brucei. Pathogens 7:33. doi: 10.3390/pathogens7020033 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203. Cárdenas C, Miller RA, Smith I, Bui T, Molgó J, Müller M, Vais H, Cheung K-H, Yang J, Parker I, Thompson CB, Birnbaum MJ, Hallows KR, Foskett JK. 2010. Essential regulation of cell bioenergetics by constitutive InsP3 receptor Ca2+ transfer to mitochondria. Cell 142:270–283. doi: 10.1016/j.cell.2010.06.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204. Cunningham KW, Fink GR. 1996. Calcineurin inhibits VCX1-dependent H+/Ca2+ exchange and induces Ca2+ ATPases in Saccharomyces cerevisiae. Mol Cell Biol 16:2226–2237. doi: 10.1128/MCB.16.5.2226 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205. Lander N, Chiurillo MA, Docampo R. 2021. Signaling pathways involved in environmental sensing in Trypanosoma cruzi. Mol Microbiol 115:819–828. doi: 10.1111/mmi.14621 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206. Moreno SN, Silva J, Vercesi AE, Docampo R. 1994. Cytosolic-free calcium elevation in Trypanosoma cruzi is required for cell invasion. J Exp Med 180:1535–1540. doi: 10.1084/jem.180.4.1535 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207. Neira I, Ferreira AT, Yoshida N. 2002. Activation of distinct signal transduction pathways in Trypanosoma cruzi isolates with differential capacity to invade host cells. Int J Parasitol 32:405–414. doi: 10.1016/s0020-7519(02)00004-8 [DOI] [PubMed] [Google Scholar]
- 208. Schenkman S, Robbins ES, Nussenzweig V. 1991. Attachment of Trypanosoma cruzi to mammalian cells requires parasite energy, and invasion can be independent of the target cell cytoskeleton. Infect Immun 59:645–654. doi: 10.1128/iai.59.2.645-654.1991 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209. Kollien AH, Schaub GA. 2000. The development of Trypanosoma cruzi in triatominae. Parasitol Today 16:381–387. doi: 10.1016/s0169-4758(00)01724-5 [DOI] [PubMed] [Google Scholar]
- 210. Lang F. 2007. Mechanisms and significance of cell volume regulation. J Am Coll Nutr 26:613S–623S. doi: 10.1080/07315724.2007.10719667 [DOI] [PubMed] [Google Scholar]
- 211. Dave N, Cetiner U, Arroyo D, Fonbuena J, Tiwari M, Barrera P, Lander N, Anishkin A, Sukharev S, Jimenez V. 2021. A novel mechanosensitive channel controls osmoregulation, differentiation, and infectivity in Trypanosoma cruzi . Elife 10. doi: 10.7554/eLife.67449 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212. Chiurillo MA, Carlson J, Bertolini MS, Raja A, Lander N, Weiss LM. 2023. Dual localization of receptor-type adenylate cyclases and cAMP response protein 3 unveils the presence of two putative signaling microdomains in Trypanosoma cruzi. mBio 14:e0106423. doi: 10.1128/mbio.01064-23 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213. Schoijet AC, Miranda K, Medeiros LCS, de Souza W, Flawiá MM, Torres HN, Pignataro OP, Docampo R, Alonso GD. 2011. Defining the role of a FYVE domain in the localization and activity of a cAMP phosphodiesterase implicated in osmoregulation in Trypanosoma cruzi. Mol Microbiol 79:50–62. doi: 10.1111/j.1365-2958.2010.07429.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214. LeFurgey A, Ingram P, Blum JJ. 2001. Compartmental responses to acute osmotic stress in Leishmania major result in rapid loss of Na+ and Cl. Comp Biochem Physiol A Mol Integr Physiol 128:385–394. doi: 10.1016/s1095-6433(00)00319-6 [DOI] [PubMed] [Google Scholar]
- 215. Docampo R, Jimenez V, Lander N, Li ZH, Niyogi S. 2013. New insights into roles of acidocalcisomes and contractile vacuole complex in osmoregulation in protists. Int Rev Cell Mol Biol 305:69–113. doi: 10.1016/B978-0-12-407695-2.00002-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216. Suess PM, Gomer RH. 2016. Extracellular polyphosphate inhibits proliferation in an autocrine negative feedback loop in Dictyostelium discoideum. J Biol Chem 291:20260–20269. doi: 10.1074/jbc.M116.737825 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217. Smith SA, Mutch NJ, Baskar D, Rohloff P, Docampo R, Morrissey JH. 2006. Polyphosphate modulates blood coagulation and fibrinolysis. Proc Natl Acad Sci U S A 103:903–908. doi: 10.1073/pnas.0507195103 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218. Smith SA, Choi SH, Davis-Harrison R, Huyck J, Boettcher J, Rienstra CM, Morrissey JH. 2010. Polyphosphate exerts differential effects on blood clotting, depending on polymer size. Blood 116:4353–4359. doi: 10.1182/blood-2010-01-266791 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219. Choi SH, Smith SA, Morrissey JH. 2015. Polyphosphate accelerates factor V activation by factor XIa. Thromb Haemost 113:599–604. doi: 10.1160/TH14-06-0515 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220. Choi SH, Smith SA, Morrissey JH. 2011. Polyphosphate is a cofactor for the activation of factor XI by thrombin. Blood 118:6963–6970. doi: 10.1182/blood-2011-07-368811 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221. Müller F, Mutch NJ, Schenk WA, Smith SA, Esterl L, Spronk HM, Schmidbauer S, Gahl WA, Morrissey JH, Renné T. 2009. Platelet polyphosphates are proinflammatory and procoagulant mediators in vivo. Cell 139:1143–1156. doi: 10.1016/j.cell.2009.11.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222. Wat JM, Foley JH, Krisinger MJ, Ocariza LM, Lei V, Wasney GA, Lameignere E, Strynadka NC, Smith SA, Morrissey JH, Conway EM. 2014. Polyphosphate suppresses complement via the terminal pathway. Blood 123:768–776. doi: 10.1182/blood-2013-07-515726 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223. Smith SA, Morrissey JH. 2008. Polyphosphate enhances fibrin clot structure. Blood 112:2810–2816. doi: 10.1182/blood-2008-03-145755 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224. Wilkinson SR, Prathalingam SR, Taylor MC, Ahmed A, Horn D, Kelly JM. 2006. Functional characterisation of the iron superoxide dismutase gene repertoire in Trypanosoma brucei. Free Radic Biol Med 40:198–209. doi: 10.1016/j.freeradbiomed.2005.06.022 [DOI] [PubMed] [Google Scholar]
- 225. Ramakrishnan S, Unger LM, Baptista RP, Cruz-Bustos T, Docampo R, Almeida IC. 2021. Deletion of a Golgi protein in Trypanosoma cruzi reveals a critical role for Mn2+ in protein glycosylation needed for host cell invasion and intracellular replication. PLoS Pathog 17:e1009399. doi: 10.1371/journal.ppat.1009399 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226. Jomova K, Makova M, Alomar SY, Alwasel SH, Nepovimova E, Kuca K, Rhodes CJ, Valko M. 2022. Essential metals in health and disease. Chem Biol Interact 367:110173. doi: 10.1016/j.cbi.2022.110173 [DOI] [PubMed] [Google Scholar]
- 227. Zhao FJ, Tang Z, Song JJ, Huang XY, Wang P. 2022. Toxic metals and metalloids: uptake, transport, detoxification, phytoremediation, and crop improvement for safer food. Molecular Plant 15:27–44. doi: 10.1016/j.molp.2021.09.016 [DOI] [PubMed] [Google Scholar]
- 228. Li FJ, He CY. 2014. Acidocalcisome is required for autophagy in Trypanosoma brucei. Autophagy 10:1978–1988. doi: 10.4161/auto.36183 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229. Docampo R, Moreno SNJ. 2008. The acidocalcisome as a target for chemotherapeutic agents in protozoan parasites. Curr Pharm Des 14:882–888. doi: 10.2174/138161208784041079 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230. Szabo CM, Oldfield E. 2001. An investigation of bisphosphonate inhibition of a vacuolar proton-pumping pyrophosphatase. Biochem Biophys Res Commun 287:468–473. doi: 10.1006/bbrc.2001.5617 [DOI] [PubMed] [Google Scholar]
- 231. Gordon-Weeks R, Parmar S, Davies TG, Leigh RA. 1999. Structural aspects of the effectiveness of bisphosphonates as competitive inhibitors of the plant vacuolar proton-pumping pyrophosphatase. Biochem J 337 (Pt 3):373–377. [PMC free article] [PubMed] [Google Scholar]
- 232. Zhen RG, Baykov AA, Bakuleva NP, Rea PA. 1994. Aminomethylenediphosphonate: a potent type-specific inhibitor of both plant and phototrophic bacterial H+-pyrophosphatases. Plant Physiol 104:153–159. doi: 10.1104/pp.104.1.153 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233. Drozdowicz YM, Shaw M, Nishi M, Striepen B, Liwinski HA, Roos DS, Rea PA. 2003. Isolation and characterization of TgVP1, a type I vacuolar H+-translocating pyrophosphatase from Toxoplasma gondii. the dynamics of its subcellular localization and the cellular effects of a diphosphonate inhibitor. J Biol Chem 278:1075–1085. doi: 10.1074/jbc.M209436200 [DOI] [PubMed] [Google Scholar]
- 234. Mathis AM, Holman JL, Sturk LM, Ismail MA, Boykin DW, Tidwell RR, Hall JE. 2006. Accumulation and intracellular distribution of antitrypanosomal diamidine compounds DB75 and DB820 in African trypanosomes. Antimicrob Agents Chemother 50:2185–2191. doi: 10.1128/AAC.00192-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235. Morrissey JH, Smith SA. 2015. Polyphosphate as modulator of hemostasis, thrombosis, and inflammation. J Thromb Haemost 13 Suppl 1:S92–7. doi: 10.1111/jth.12896 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236. Beyersmann D, Haase H. 2001. Functions of zinc in signaling, proliferation and differentiation of mammalian cells. Biometals 14:331–341. doi: 10.1023/a:1012905406548 [DOI] [PubMed] [Google Scholar]









