Abstract
Bacterial division initiates at the site of a contractile Z-ring composed of polymerized FtsZ. The location of the Z-ring in the cell is controlled by a system of three mutually antagonistic proteins, MinC, MinD, and MinE. Plastid division is also known to be dependent on homologs of these proteins, derived from the ancestral cyanobacterial endosymbiont that gave rise to plastids. In contrast, the mitochondria of model systems such as Saccharomyces cerevisiae, mammals, and Arabidopsis thaliana seem to have replaced the ancestral α-proteobacterial Min-based division machinery with host-derived dynamin-related proteins that form outer contractile rings. Here, we show that the mitochondrial division system of these model organisms is the exception, rather than the rule, for eukaryotes. We describe endosymbiont-derived, bacterial-like division systems comprising FtsZ and Min proteins in diverse less-studied eukaryote protistan lineages, including jakobid and heterolobosean excavates, a malawimonad, stramenopiles, amoebozoans, a breviate, and an apusomonad. For two of these taxa, the amoebozoan Dictyostelium purpureum and the jakobid Andalucia incarcerata, we confirm a mitochondrial localization of these proteins by their heterologous expression in Saccharomyces cerevisiae. The discovery of a proteobacterial-like division system in mitochondria of diverse eukaryotic lineages suggests that it was the ancestral feature of all eukaryotic mitochondria and has been supplanted by a host-derived system multiple times in distinct eukaryote lineages.
Keywords: mitochondria, mitochondrial division, Min proteins, MinCDE, mitochondrial fission
During bacterial division, septum formation is mediated by the Z-ring, a contractile ring structure made up of the polymerized tubulin homolog FtsZ (reviewed in refs. 1 and 2). The site at which FtsZ polymerizes is determined by the Min system (reviewed in refs. 1 and 3–5), comprising the three septum site-determining proteins MinC, MinD, and MinE. ATP-bound, dimerized MinD binds the inner cell membrane at the poles of the cell, forming aggregates. These MinD aggregates bind and activate dimerized MinC (6), which then inhibits local FtsZ polymerization (Fig. 1A). Concomitantly, dimerized MinE forms a spiral ring whose constant polymerization and depolymerization causes it to oscillate across the cell (7, 8). Where MinE comes into contact with MinD, it causes the release of ATP and the subsequent liberation of MinD from the membrane (9, 10). In this way, MinD and MinC cannot inhibit FtsZ polymerization near the midpoint of the cell. The polymerizing Z-ring is stabilized and tethered to the membrane by FtsA, ZipA, and the nonessential ZapA and (in some organisms) ZapB (11–15). Maturation of the Z-ring into a complete septal ring continues with the subsequent recruitment by FtsA of further components of the divisome (i.e., FtsB, FtsE, FtsI/PBP3, FtsK, FtsL, FtsN, FtsQ, FtsW, and FtsX), which proceed to stabilize FtsZ and contribute to peptidoglycan synthesis (reviewed in refs. 3, 4, 16 and 17) (Fig. 1B) before Z-ring constriction and the completion of septum formation (Fig. 1C).
Fig. 1.
Partial schematic overview of division machinery in E. coli. (A) Roles of Min proteins during FtsZ polymerization. (B) Subsequent recruitment of early and late stage proteins involved in Z-ring stabilization and attachment to the cell membrane. (C) Overview of septation initiation at the cell level. Dark-blue rectangles, FtsZ; dark-green circles, MinD; light-blue shapes, late-stage cell-division proteins; light-green circles, MinC; magenta circles, MinE; red shapes, early-stage cell-division proteins. For the sake of clarity, not all proteins known to localize to the mid-cell during division are shown. In particular, this schematic focuses on proteins known to localize to the cytoplasmic membrane and excludes most proteins localizing primarily to the peptidoglycan layer and the outer membrane. Based on reviews in refs. 1, 16, and 17.
Plastids are known to possess FtsZ (18–20), MinD (21, 22), MinE (21, 23), and, in some cases, MinC (24) homologs of cyanobacterial endosymbiotic origin; in some cases, the latter are encoded on the plastid genome (21, 25). In contrast, only two examples of putative mitochondrial Min proteins have been reported, in the stramenopiles Nannochloropsis oceanica and Ectocarpus siliculosus (26). Indeed, although eukaryotic mitochondria are derived from an α-proteobacterial endosymbiont, the ancestral bacterial division machinery has been partly or wholly replaced by eukaryote-specific proteins in model system eukaryotes where mitochondrial division has been studied. Whereas Amoebozoa (27), stramenopiles (28, 29), and the red alga Cyanidioschyzon merolae (30, 31) have retained experimentally confirmed mitochondrial FtsZ, animals and fungi (opisthokonts) and plants examined to date lack this protein. In the latter taxa, an outer contractile ring is instead formed by Dnm1p/Drp1, a eukaryote-specific dynamin GTPase (32–34). This protein is implicated in mitochondrial division in organisms across the eukaryotic tree, including Arabidopsis thaliana (35–37), the parabasalid Trichomonas vaginalis (38), Dictyostelium discoideum (39), and C. merolae (40), suggesting that the outer contractile ring is a widespread eukaryotic feature. In T. vaginalis and A. thaliana, the nature of the inner contractile ring is not yet understood, although the presence of two Dnm1/Drp1 homologs in A. thaliana (37) raises the possibility that they form an outer and an inner contractile ring, respectively. Recent work (41) reconstructing the evolution of eukaryotic dynamins suggests that the ancestral mitochondrial dynamin was a bifunctional protein that also mediated vesicle scission. This protein underwent duplication events, followed by subfunctionalization, independently in at least three lineages (opisthokonts, land plants, and alveolates), but the ancestral bifunctional form seems to have been retained in amoebozoa such as D. discoideum, the red alga C. merolae, and stramenopiles (and possibly additional eukaryotes that have currently less well-characterized dynamins). The distribution of ancestral-like bifunctional mitochondrial/vesicle fission dynamins thus seems to mirror that of mitochondrial FtsZ (41).
Here, we hypothesize that the complete loss of the α-proteobacterial division system is the exception, rather than the rule, for eukaryotes. We show that mitochondria-targeted homologs of bacterial Min proteins are patchily but widely distributed among diverse eukaryote lineages; and we further demonstrate that Min proteins from two of these lineages, the amoebozoan Dictyostelium purpureum and the jakobid excavate Andalucia incarcerata, localize to mitochondria when expressed in yeast.
Materials and Methods
Database Searches.
Publicly available databases and sequencing projects were searched using the Basic Local Alignment Search Tools (BLAST) blastp and tblastn (42). A large number of databases containing eukaryotic sequences were screened with these tools using query sequences from D. purpureum (XP_003286111, XP_003292258, XP_003293637, XP_642499), E. siliculosus (CBJ32744, CBJ31561, CBJ28079, CBJ48312) A. incarcerata, Pseudomonas fluorescens (AEV64338, AEV64339, AEV64340, AEV64767), and Anabaena sp. 90 (YP_006998153, AFW94434, YP_006996248, YP_006996249). The databases searched included the Nucleotide collection (nr/nt), National Center for Biotechnology Information (NCBI) Genomes, Whole-Genome Shotgun contigs, Expressed Sequence Tags, High-throughput Genomic Sequences and Transcriptome Shotgun Assembly divisions of GenBank (43) (last accessed February 9, 2015); the Broad Institute project databases (44) (accessed April 23, 2014); the Joint Genome Institute (JGI) genome databases (45, 46) (last accessed February 9, 2015); dictyBase, 2013 release (47–49); the EnsemblProtists database (50) (last accessed February 9, 2015); the Eukaryotic Pathogen Database Resources (EuPathDB) (51) (last accessed February 9, 2015); and the Marine Microbial Eukaryote Transcriptome Sequencing Project (52) (accessed June 3, 2014), via the Community cyberinfrastructure for Advanced Microbial Ecology Research and Analysis (CAMERA) portal (53) (for a full list of sequences identified, see Table S1 and Dataset S1). In addition, we searched our own unpublished genome or transcriptome assemblies from several protist taxa of key evolutionary interest: two jakobids (A. incarcerata and Andalucia godoyi), the heterolobosean Pharyngomonas kirbyi, and Malawimonas californiana. Potential homologs identified were screened manually to exclude contaminants from bacterial or other eukaryotic sources, by searching for introns and excluding sequences with a notably high degree of similarity to bacterial or distantly related eukaryotic homologs. Subcellular localization and targeting peptides were predicted using TargetP, using “plant” parameters for plastid-bearing taxa and “nonplant” parameters for taxa lacking plastids (54, 55).
Sequence Generation.
P. kirbyi strain AS12B (56, 57) was cultivated at 37 °C in 10% (wt/vol) salt medium (NaCl 1.6 M, KCl 34.0 mM, MgCl2 44.2 mM, CaCl2 4.0 mM, MgSO4 4.5 mM) supplemented with Citrobacter sp. as a food source before RNA isolation. RNA was extracted using TRIzol (Life Technologies) following the manufacturer’s instructions and stored at −80 °C. The RNA sample was treated with Turbo DNase (Life Technologies) before conversion to cDNA using the GeneRacer kit with SuperScript III reverse transcriptase (Life Technologies) and stored at −20 °C. Primers were designed to amplify genes of interest using available sequences. Primer sequences were as follows: MinCF, 5′-ATGTCACGTCGATGGTTAGT-3′; MinCR, 5′-TAATACAAAAAAAAAACA-3′; MinDF, 5′-ATGTATCGATCAACGAGTTC-3′; and MinDR, 5′-TTAGTTCCTGCTAAATAATC-3′. PCR reactions were done using the Phusion high-fidelity DNA polymerase (New England BioLabs) where the initial denaturation at 98 °C for 30 s was followed by 30 cycles of DNA denaturation at 98 °C for 10 s, primer annealing at 40 °C for 30 s, and strand elongation at 72 °C for 60 s, with a final extension at 72 °C for 10 s. PCR products were purified by gel extraction using the Nucleospin Extract II kit (Macherey-Nagel) and were directly sequenced using the PCR primers.
Phylogenetic Analyses.
For each protein, alignments were generated from datasets including all known eukaryotic homologs and bacterial homologs harvested from NCBI using MUSCLE v.3.8.31 (58) or MAFFT-L-INSI v7.149b (59–61), and trimmed using BMGE 1.1 (62) (-m BLOSUM30; all other parameters default). Preliminary phylogenies were generated using FastTree, and datasets were manually refined. Twenty independent Maximum Likelihood (ML) tree estimates and 200 bootstrap replicates were generated using RAxML v.8.0.23 (63) under the PROTGAMMALG4X (64) model of amino acid substitution. Bayesian inference posterior probabilities were calculated using PhyloBayes v.3.3f (65) under the catfix C20 model of evolution. We tested whether specific phylogenetic hypotheses were rejected by the data using the approximately unbiased (AU) test implemented in CONSEL v.1.20 (66) (Table S2). Maximum-likelihood trees given specific constraints (i.e., corresponding to specific hypotheses) were generated using RAxML. In addition, the 200 trees from bootstrap replicates were included in the hypothesis-testing analyses performed with CONSEL.
Yeast Culture, Transformation, and Microscopy.
Saccharomyces cerevisiae strain YPH499 was grown at 30 °C on YPD medium or selective medium without uracil after lithium-acetate transformation. For ectopic expression of AiMinC, -D, and -E, the complete AiMinC, -D, and -E ORFs were amplified by PCR from A. incarcerata cDNA. For ectopic expression of DpMinC, -D, and -E, the complete DpMinC, -D, and -E ORFs were amplified by PCR from synthesized DNA fragments, containing Escherichia coli codon-optimized sequences. The resulting PCR products were cloned separately into pUG35 using XbaI/ClaI restriction sites (AiMinD and DpMinC and -E) or BamHI/HindIII restriction sites (AiMinC and -E and DpMinD), allowing the expression of GFP on the C-terminus of each protein. For fluorescence microscopy, cells were incubated with MitoTracker Red CMXRos (1:10,000) for 10 min, washed once in PBS, and mounted in 2% low-melting agarose. Cells were viewed using an Olympus IX81 microscope and a Hamamatsu Orca-AG digital camera using the cell^R imaging program at 100× magnification.
Results and Discussion
We identified sequences encoding at least one Min protein from a number of eukaryotic taxa (Fig. 2 and Table S1), including ancestrally plastid-lacking lineages such as the apusomonad Thecamonas trahens, the breviate Pygsuia biforma, the jakobid excavates A. godoyi and A. incarcerata, the malawimonad M. californiana, and several amoebozoan lineages, such as D. purpureum. Three previously reported FtsZ sequences identified in haptophytes (Gephyrocapsa oceanica and Pleurochrysis carterae) and a glaucophyte (Cyanophora paradoxa) (29) were excluded as probable α-proteobacterial contaminants, based on their position in preliminary phylogenies, their high degree of similarity to α-proteobacterial sequences, and, in the case of C. paradoxa, our inability to recover the reported mitochondrial FtsZ sequence from the genome sequence (67). All complete genomes encoding at least one Min protein also encoded at least one FtsZ homolog; however, the reverse was not true. Min proteins were retained not only in lineages with typical aerobic mitochondria, but also in lineages possessing mitochondrion-related organelles (MROs) such as A. incarcerata (68) and P. biforma (69).
Fig. 2.
Presence and absence of bacterial Min proteins and FtsZ in selected eukaryotic taxa. Blue, predicted mitochondrial proteins; gray, no protein found encoded in complete genome data; green, predicted plastid proteins; ?, no protein found encoded in transcriptome or incomplete genome data; *, chromatophore protein; †, predicted pseudogene; ‡, with the exception of Physcomitrella patens. Boxes shaded half blue and half green represent multiple paralogs, predicted to be mitochondrial and plastid, respectively. In cases where only a transcriptome or incomplete genome is available, it should be noted that the presence of a plastid protein does not exclude the possibility of one or more mitochondrial paralogs also being present, and vice versa. Eukaryotic taxa possessing predicted mitochondrial Min proteins are shaded in blue. Mitochondrial or plastid predictions are based on phylogenetic affinity with previously localized proteins, predicted subcellular localization, and localization in yeast (A. incarcerata, D. discoideum). Black circles indicate taxa in which reticulate mitochondria have previously been described; gray circles indicate groups for which reticulate mitochondria have been described in at least one member; black-bordered white circles indicate taxa in which only single or unbranched mitochondria have been described. The schematic phylogeny reflects the current understanding of relationships based on multiple phylogenomic analyses. For a more complete table, see Table S1.
Most of these Min and FtsZ homologs possess predicted mitochondrial targeting peptides (Table S1). To confirm these predictions, we expressed GFP-tagged homologs of Min proteins in S. cerevisiae, in conjunction with the mitochondrial stain MitoTracker Red CMXRos (Fig. 3). We chose Min proteins from two representative taxa lacking plastids: the amoebozoan D. purpureum (Fig. 3A) and the jakobid excavate A. incarcerata (Fig. 3B). In both cases, the GFP signal colocalized with the Mitotracker signal, supporting the predicted targeting of A. incarcerata and D. discoideum Min proteins to the inside of the mitochondria.
Fig. 3.
Min proteins from D. purpureum (A) and A. incarcerata (B) expressed in S. cerevisiae to confirm predicted mitochondrial targeting. Differential interference contrast (DIC) images of S. cerevisiae cells expressing Min fusion proteins (Left); in green, MinC, MinD, or MinE expressed with the C-terminal GFP tag in S. cerevisiae; in red, mitochondria labeled with MitoTracker Red CMXRos (Mito); merged images (Merge) show mitochondrial localization of all Min proteins. (Scale bars: 5 µm.)
Single-protein phylogenies of MinC, -D, -E, and FtsZ recover all predicted mitochondrial homologs as well-resolved clades, distinct from known and predicted plastid sequences (Figs. 4 and 5 and Figs. S1 and S2). For MinD (Fig. 4) and FtsZ (Fig. 5 and Fig. S3), the hypothesis that the plastid and mitochondrial homologs group in a monophyletic clade was rejected by AU tests (Table S2); however, this hypothesis could not be rejected for the more divergent MinC (Fig. S1) and MinE (Fig. S2). In all three Min phylogenies, mitochondrial homologs emerged within proteobacterial sequences although, because the resolution within that clade was too poor to identify the closest homologs (Fig. 4 and Figs. S1 and S2), we cannot exclude the possibility that these proteins originate from a group other than the α-proteobacteria.
Fig. 4.
Unrooted maximum likelihood (ML) tree of MinD sequences. Phylogenetic analyses were performed on 328 sequences and 226 sites, using RAxML and PhyloBayes. Bootstrap support values greater than 50% and posterior probabilities greater than 0.5 are shown. Branches with 100% bootstrap support and posterior probability of 1.0 are indicated by black circles. Eukaryotes are shaded blue, cyanobacteria green, proteobacteria orange, and α-proteobacteria magenta.
Fig. 5.
Unrooted maximum likelihood (ML) tree of FtsZ sequences. Phylogenetic analyses were performed on 327 sequences and 257 sites, using RAxML and PhyloBayes. Bootstrap support values greater than 50% and posterior probabilities greater than 0.5 are shown. Branches with 100% bootstrap support and posterior probability of 1.0 are indicated by black circles. Eukaryotes are shaded blue, cyanobacteria green, proteobacteria orange, and α-proteobacteria magenta. Eukaryotic paralogs lacking the variable C-terminal spacer region are indicated by stars whereas those with incomplete sequence at the C-terminus are indicated by question marks. The exception to this pattern is a Corethron hystrix sequence that, despite branching with other stramenopiles in the MtFtsZ1 clade, possesses a C-terminal variable region (Fig. S3).
An early study by Miyagishima et al. (70) reported the presence of two copies of plastid-targeted FtsZ in photosynthetic eukaryotes, as well as two copies of predicted mitochondrial FtsZ in C. merolae and D. discoideum. The authors hypothesized that duplication of FtsZ occurred early during primary plastid endosymbiosis and that a similar process might also have accompanied the establishment of the protomitochondrial endosymbiont. Our broader taxonomic sampling allowed us to confirm the presence of two types of mitochondrial FtsZ homolog in the majority of the eukaryotic taxa examined. They form two distinct phylogenetic clades, each of which contains one homolog from each eukaryote. In addition, although these clades lack strong statistical support, one encompasses copies retaining a variable C-terminal spacer domain that is also found in bacterial homologs (71) whereas sequences from the other clade lack this domain (Fig. 5 and Fig. S3). A robust grouping of α-proteobacterial and both putative mitochondrial FtsZ paralogs was recovered. We subsampled these sequences, using γ-proteobacterial sequences as an outgroup, and reanalyzed them in an attempt to better resolve this clade. We also excluded amoebozoan sequences because of their unusually high AT content and long branches in preliminary trees (Fig. S3). Unfortunately, we were unable to obtain better resolution of the branching order among α-proteobacterial and eukaryotic clades. Nevertheless, the C-terminally truncated FtsZ proteins were identified only in eukaryotes, and we did not identify more than one FtsZ homolog in α-proteobacteria. We therefore conclude that the duplication event that gave rise to the FtsZ paralogs found in extant eukaryotes likely occurred early in eukaryotic evolution, rather than earlier, in the α-proteobacterial lineage that gave rise to the mitochondrion.
Altogether, these lines of evidence are consistent with the hypothesis that the nuclear-encoded mitochondrial Min and FtsZ homologs of eukaryotes originated by endosymbiotic gene transfer from the ancestral mitochondrial endosymbiont.
Although found in diverse eukaryotes, the Min proteins are sparsely distributed, a pattern that can only partly be reconciled with taxonomic representation in the available data. A striking example of gene loss is seen in the Mycetozoa (Dictyostelium spp., Acytostelium subglobosum, and Polysphondylium pallidum). Here, D. discoideum, Dictyostelium citrinum, Dictyostelium intermedium, and Dictyostelium firmibasis have retained only FtsZ whereas their sister taxon D. purpureum and the more basal taxa P. pallidum, Polysphondylium violaceum, and A. subglobosum have additionally maintained all three Min proteins. Meanwhile, the yet more distantly related Dictyostelium fasciculatum (72, 73) seems to have independently lost the Min proteins and, like D. discoideum, only possesses FtsZ. This overall pattern raises the question of why Min proteins were retained in some taxa, yet lost in others. No correlation was found with mitochondrial cristae morphology, because Min proteins were found in organisms possessing discoid (e.g., P. kirbyi) (56) or tubular (e.g., A. godoyi) (74) cristae, as well as in lineages with MROs that apparently lack cristae entirely (e.g., A. incarcerata and P. biforma) (68, 69). Nor is there any obvious difference in either overall mitochondrial morphology or lifestyle between lineages that possess Min proteins and lineages that do not. Kiefel et al. (29) have raised the possibility that FtsZ is lost in lineages with reticulate mitochondria, and thus the placement of the division site may not affect mitochondrial function. This hypothesis remains a plausible explanation that might apply to Min proteins; A. godoyi and P. biforma each possess a single mitochondrion or mitochondrion-related organelle (74, 75), and T. trahens is predicted to have discrete, nonbranching mitochondria based on 3D reconstructions (76) (Fig. 2). Meanwhile, a number of the lineages lacking Min proteins are known to possess reticulate mitochondria in at least one tissue and during at least one life stage, including opisthokonts (77), plants (78, 79), the euglenozoan Euglena gracilis (80), and apicomplexa (81, 82) (Fig. 2). One exception seems to be Phytophthora cinnamomi, an organism described in the literature as having 3–4 reticulate mitochondria per cell (83), and in which we found Min homologs. Unfortunately, there are relatively few taxa in our survey for which detailed microscopy data are available that would permit conclusions to be drawn about the 3D structure of their mitochondria. Furthermore, many organisms known to possess reticulate mitochondria may also possess unbranched mitochondria in some tissues, during some parts of their life cycle (79), or alongside reticulate mitochondria, as in P. cinnamomi (83). It is therefore also possible that the presence or absence of Min proteins reflects some unknown transient mitochondrial morphological feature specific to replication. Clearly, genetic and functional studies of mitochondrial Min systems are greatly needed to understand their precise roles.
Further questions are raised by the apparent absence of homologs of all other components of the bacterial divisome from the surveyed eukaryotes, including ZipA, ZapA, FtsA, FtsB, FtsE, FtsI, FtsK, FtsL, FtsN, FtsQ, FtsW, and FtsX. Searches of databases using α-proteobacterial and E. coli homologs of these proteins as queries yielded no candidate homologs. The bacterial divisome components recruited late in the division process (FtsB, FtsE, FtsI, FtsL, FtsN, FtsQ, FtsW, and FtsX) are primarily involved in facilitating peptidoglycan synthesis, and so their apparent absence is perhaps not surprising, given the lack of a peptidoglycan wall in any mitochondria. However, it is not clear how the Z-ring remains stabilized and anchored to the membrane in the absence of FtsA, ZipA, or ZapA. ZED, a coiled-coil domain protein with 25.8% sequence identity to ZapA, is reported to be involved in mitochondrial Z-ring formation in the red alga C. merolae (84). However, we were unable to identify any homologs of this protein in other eukaryotes. The two distinct FtsZ paralogs may form an alternating copolymer that forms the Z-ring; or the Z-ring might be composed of a single paralog whereas the second paralog might instead be involved in attachment of the Z-ring to the membrane. In either case, the anchoring mechanism of FtsZ remains a mystery.
Recent work (85) implicates the endoplasmic reticulum (ER) in the control of the mitochondrial division site location and subsequent Dnm1p recruitment in yeast. This type of external division site control contrasts with that of the Min protein system, which regulates division site location from the mitochondrial matrix. The contrast between these control mechanisms raise the questions of when the role of the ER in mitochondrial division may have emerged; whether any taxa possess both Min proteins and Dnm1p/Drp1; and how these organisms (if they exist) recruit Dnm1p/Drp1 in the absence of ER-mediated division site control. Therefore, an important avenue of further study is the taxonomic distribution of mitochondrial Dnm1p/Drp1 and its functional interplay with FtsZ. Study of this distribution is hampered by the fact that multiple paralogs of dynamins have different functions within eukaryotic cells (41), including vesicular trafficking in yeast (86), and unknown functions in less-studied organisms such as T. vaginalis (38). These proteins lack N-terminal targeting peptides, and so, in the absence of localization data, a mitochondrial function cannot clearly be ascribed to any one of them based on sequence data alone. In any case, investigations into the molecular mechanisms governing the coordination of the various kinds of inner and outer contractile rings are critically needed in diverse eukaryote lineages to fully understand what are features of the division system of the last eukaryotic common ancestor and what are more recent lineage-specific innovations.
Supplementary Material
Acknowledgments
We thank Dr. Michael W. Gray for planting the seeds of the collaboration leading to this paper. We thank the Acytostelium genome consortium for kindly providing A. subglobosum gene sequences. M.M.L. thanks Yana Eglit for help in taming Adobe Illustrator. M.M.L. was supported by the National Research Fund, Luxembourg (FNR), and by the Nova Scotia Health Research Foundation (NSHRF). L.E. was supported by a Centre for Comparative Genomics and Evolutionary Bioinformatrics postdoctoral fellowship from the Tula Foundation. T.H. was supported by the Natural Sciences and Engineering Research Council of Canada. This work was supported by Regional Partnerships Program Grant FRN 62809 from the Canadian Institutes of Health Research and the NSHRF (to A.J.R.), Czech Science Foundation Grant 13-24983S (to M.E.), Czech Science Foundation Grant 13-29423S (to P.D.), and a European Regional Development Fund award to the Biomedicine Center of the Academy of Sciences and Charles University (CZ.1.05/1.1.00/02.0109).
Footnotes
The authors declare no conflict of interest.
This paper results from the Arthur M. Sackler Colloquium of the National Academy of Sciences, “Symbioses Becoming Permanent: The Origins and Evolutionary Trajectories of Organelles,” held October 15–17, 2014, at the Arnold and Mabel Beckman Center of the National Academies of Sciences and Engineering in Irvine, CA. The complete program and video recordings of most presentations are available on the NAS website at www.nasonline.org/Symbioses.
This article is a PNAS Direct Submission. P.J.K. is a guest editor invited by the Editorial Board.
Data deposition: The sequences reported in this paper have been deposited in the GenBank database (accession nos. KP271960–KP271964, KP324909–KP324912, KP258196–KP258204, and KP738110).
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1421392112/-/DCSupplemental.
References
- 1.de Boer PA. Advances in understanding E. coli cell fission. Curr Opin Microbiol. 2010;13(6):730–737. doi: 10.1016/j.mib.2010.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Meier EL, Goley ED. Form and function of the bacterial cytokinetic ring. Curr Opin Cell Biol. 2014;26:19–27. doi: 10.1016/j.ceb.2013.08.006. [DOI] [PubMed] [Google Scholar]
- 3.Natale P, Pazos M, Vicente M. The Escherichia coli divisome: Born to divide. Environ Microbiol. 2013;15(12):3169–3182. doi: 10.1111/1462-2920.12227. [DOI] [PubMed] [Google Scholar]
- 4.Lutkenhaus J, Pichoff S, Du S. Bacterial cytokinesis: From Z ring to divisome. Cytoskeleton (Hoboken) 2012;69(10):778–790. doi: 10.1002/cm.21054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Lutkenhaus J. Assembly dynamics of the bacterial MinCDE system and spatial regulation of the Z ring. Annu Rev Biochem. 2007;76:539–562. doi: 10.1146/annurev.biochem.75.103004.142652. [DOI] [PubMed] [Google Scholar]
- 6.Ghosal D, Trambaiolo D, Amos LA, Löwe J. MinCD cell division proteins form alternating copolymeric cytomotive filaments. Nat Commun. 2014;5:5341. doi: 10.1038/ncomms6341. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Hale CA, Meinhardt H, de Boer PA. Dynamic localization cycle of the cell division regulator MinE in Escherichia coli. EMBO J. 2001;20(7):1563–1572. doi: 10.1093/emboj/20.7.1563. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Loose M, Fischer-Friedrich E, Ries J, Kruse K, Schwille P. Spatial regulators for bacterial cell division self-organize into surface waves in vitro. Science. 2008;320(5877):789–792. doi: 10.1126/science.1154413. [DOI] [PubMed] [Google Scholar]
- 9.Park KT, Wu W, Lovell S, Lutkenhaus J. Mechanism of the asymmetric activation of the MinD ATPase by MinE. Mol Microbiol. 2012;85(2):271–281. doi: 10.1111/j.1365-2958.2012.08110.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Hu Z, Lutkenhaus J. Topological regulation of cell division in E. coli. spatiotemporal oscillation of MinD requires stimulation of its ATPase by MinE and phospholipid. Mol Cell. 2001;7(6):1337–1343. doi: 10.1016/s1097-2765(01)00273-8. [DOI] [PubMed] [Google Scholar]
- 11.Hale CA, de Boer PA. Direct binding of FtsZ to ZipA, an essential component of the septal ring structure that mediates cell division in E. coli. Cell. 1997;88(2):175–185. doi: 10.1016/s0092-8674(00)81838-3. [DOI] [PubMed] [Google Scholar]
- 12.Wang X, Huang J, Mukherjee A, Cao C, Lutkenhaus J. Analysis of the interaction of FtsZ with itself, GTP, and FtsA. J Bacteriol. 1997;179(17):5551–5559. doi: 10.1128/jb.179.17.5551-5559.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Pichoff S, Lutkenhaus J. Unique and overlapping roles for ZipA and FtsA in septal ring assembly in Escherichia coli. EMBO J. 2002;21(4):685–693. doi: 10.1093/emboj/21.4.685. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Hale CA, de Boer PA. Recruitment of ZipA to the septal ring of Escherichia coli is dependent on FtsZ and independent of FtsA. J Bacteriol. 1999;181(1):167–176. doi: 10.1128/jb.181.1.167-176.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Galli E, Gerdes K. Spatial resolution of two bacterial cell division proteins: ZapA recruits ZapB to the inner face of the Z-ring. Mol Microbiol. 2010;76(6):1514–1526. doi: 10.1111/j.1365-2958.2010.07183.x. [DOI] [PubMed] [Google Scholar]
- 16.Egan AJ, Vollmer W. The physiology of bacterial cell division. Ann N Y Acad Sci. 2013;1277:8–28. doi: 10.1111/j.1749-6632.2012.06818.x. [DOI] [PubMed] [Google Scholar]
- 17.den Blaauwen T, Andreu JM, Monasterio O. Bacterial cell division proteins as antibiotic targets. Bioorg Chem. 2014;55:27–38. doi: 10.1016/j.bioorg.2014.03.007. [DOI] [PubMed] [Google Scholar]
- 18.Fraunholz MJ, Moerschel E, Maier UG. The chloroplast division protein FtsZ is encoded by a nucleomorph gene in cryptomonads. Mol Gen Genet. 1998;260(2-3):207–211. doi: 10.1007/s004380050887. [DOI] [PubMed] [Google Scholar]
- 19.Sato M, Nishikawa T, Yamazaki T, Kawano S. Isolation of the plastid ftsZ gene from Cyanophora paradoxa (Glaucocystophyceae, Glaucocystophyta) Phycol Res. 2005;53:93–96. [Google Scholar]
- 20.Osteryoung KW, Vierling E. Conserved cell and organelle division. Nature. 1995;376(6540):473–474. doi: 10.1038/376473b0. [DOI] [PubMed] [Google Scholar]
- 21.Wakasugi T, et al. Complete nucleotide sequence of the chloroplast genome from the green alga Chlorella vulgaris: The existence of genes possibly involved in chloroplast division. Proc Natl Acad Sci USA. 1997;94(11):5967–5972. doi: 10.1073/pnas.94.11.5967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Colletti KS, et al. A homologue of the bacterial cell division site-determining factor MinD mediates placement of the chloroplast division apparatus. Curr Biol. 2000;10(9):507–516. doi: 10.1016/s0960-9822(00)00466-8. [DOI] [PubMed] [Google Scholar]
- 23.Itoh R, Fujiwara M, Nagata N, Yoshida S. A chloroplast protein homologous to the eubacterial topological specificity factor minE plays a role in chloroplast division. Plant Physiol. 2001;127(4):1644–1655. [PMC free article] [PubMed] [Google Scholar]
- 24.Miyagishima SY, Kabeya Y, Sugita C, Sugita M, Fujiwara T. DipM is required for peptidoglycan hydrolysis during chloroplast division. BMC Plant Biol. 2014;14:57. doi: 10.1186/1471-2229-14-57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Douglas SE, Penny SL. The plastid genome of the cryptophyte alga, Guillardia theta: Complete sequence and conserved synteny groups confirm its common ancestry with red algae. J Mol Evol. 1999;48(2):236–244. doi: 10.1007/pl00006462. [DOI] [PubMed] [Google Scholar]
- 26.Vieler A, et al. Genome, functional gene annotation, and nuclear transformation of the heterokont oleaginous alga Nannochloropsis oceanica CCMP1779. PLoS Genet. 2012;8(11):e1003064. doi: 10.1371/journal.pgen.1003064. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Gilson PR, et al. Two Dictyostelium orthologs of the prokaryotic cell division protein FtsZ localize to mitochondria and are required for the maintenance of normal mitochondrial morphology. Eukaryot Cell. 2003;2(6):1315–1326. doi: 10.1128/EC.2.6.1315-1326.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Beech PL, et al. Mitochondrial FtsZ in a chromophyte alga. Science. 2000;287(5456):1276–1279. doi: 10.1126/science.287.5456.1276. [DOI] [PubMed] [Google Scholar]
- 29.Kiefel BR, Gilson PR, Beech PL. Diverse eukaryotes have retained mitochondrial homologues of the bacterial division protein FtsZ. Protist. 2004;155(1):105–115. doi: 10.1078/1434461000168. [DOI] [PubMed] [Google Scholar]
- 30.Takahara M, et al. A putative mitochondrial ftsZ gene is present in the unicellular primitive red alga Cyanidioschyzon merolae. Mol Gen Genet. 2000;264(4):452–460. doi: 10.1007/s004380000307. [DOI] [PubMed] [Google Scholar]
- 31.Takahara M, Kuroiwa H, Miyagishima S, Mori T, Kuroiwa T. Localization of the mitochondrial FtsZ protein in a dividing mitochondrion. Cytologia (Tokyo) 2001;66:421–425. [Google Scholar]
- 32.Bleazard W, et al. The dynamin-related GTPase Dnm1 regulates mitochondrial fission in yeast. Nat Cell Biol. 1999;1(5):298–304. doi: 10.1038/13014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Labrousse AM, Zappaterra MD, Rube DA, van der Bliek AM. C. elegans dynamin-related protein DRP-1 controls severing of the mitochondrial outer membrane. Mol Cell. 1999;4(5):815–826. doi: 10.1016/s1097-2765(00)80391-3. [DOI] [PubMed] [Google Scholar]
- 34.Smirnova E, Griparic L, Shurland DL, van der Bliek AM. Dynamin-related protein Drp1 is required for mitochondrial division in mammalian cells. Mol Biol Cell. 2001;12(8):2245–2256. doi: 10.1091/mbc.12.8.2245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Arimura S, Tsutsumi N. A dynamin-like protein (ADL2b), rather than FtsZ, is involved in Arabidopsis mitochondrial division. Proc Natl Acad Sci USA. 2002;99(8):5727–5731. doi: 10.1073/pnas.082663299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Mano S, Nakamori C, Kondo M, Hayashi M, Nishimura M. An Arabidopsis dynamin-related protein, DRP3A, controls both peroxisomal and mitochondrial division. Plant J. 2004;38(3):487–498. doi: 10.1111/j.1365-313X.2004.02063.x. [DOI] [PubMed] [Google Scholar]
- 37.Arimura S, Aida GP, Fujimoto M, Nakazono M, Tsutsumi N. Arabidopsis dynamin-like protein 2a (ADL2a), like ADL2b, is involved in plant mitochondrial division. Plant Cell Physiol. 2004;45(2):236–242. doi: 10.1093/pcp/pch024. [DOI] [PubMed] [Google Scholar]
- 38.Wexler-Cohen Y, Stevens GC, Barnoy E, van der Bliek AM, Johnson PJ. A dynamin-related protein contributes to Trichomonas vaginalis hydrogenosomal fission. FASEB J. 2014;28(3):1113–1121. doi: 10.1096/fj.13-235473. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Wienke DC, Knetsch ML, Neuhaus EM, Reedy MC, Manstein DJ. Disruption of a dynamin homologue affects endocytosis, organelle morphology, and cytokinesis in Dictyostelium discoideum. Mol Biol Cell. 1999;10(1):225–243. doi: 10.1091/mbc.10.1.225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Nishida K, et al. Dynamic recruitment of dynamin for final mitochondrial severance in a primitive red alga. Proc Natl Acad Sci USA. 2003;100(4):2146–2151. doi: 10.1073/pnas.0436886100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Purkanti R, Thattai M. Ancient dynamin segments capture early stages of host-mitochondrial integration. Proc Natl Acad Sci USA. 2015;112(9):2800–2805. doi: 10.1073/pnas.1407163112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Altschul SF, et al. Gapped BLAST and PSI-BLAST: A new generation of protein database search programs. Nucleic Acids Res. 1997;25(17):3389–3402. doi: 10.1093/nar/25.17.3389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Benson DA, et al. GenBank. Nucleic Acids Res. 2014;42(Database issue):D32–D37. doi: 10.1093/nar/gkt1030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Origins of Multicellularity Sequencing Project, Broad Institute of Harvard and MIT www.broadinstitute.org/. Accessed April 23, 2014.
- 45.Grigoriev IV, et al. The genome portal of the Department of Energy Joint Genome Institute. Nucleic Acids Res. 2012;40(Database issue):D26–D32. doi: 10.1093/nar/gkr947. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Nordberg H, et al. The genome portal of the Department of Energy Joint Genome Institute: 2014 updates. Nucleic Acids Res. 2014;42(Database issue):D26–D31. doi: 10.1093/nar/gkt1069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Kreppel L, et al. dictyBase: A new Dictyostelium discoideum genome database. Nucleic Acids Res. 2004;32(Database issue):D332–D333. doi: 10.1093/nar/gkh138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Basu S, et al. DictyBase 2013: Integrating multiple Dictyostelid species. Nucleic Acids Res. 2013;41(Database issue):D676–D683. doi: 10.1093/nar/gks1064. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Fey P, Dodson RJ, Basu S, Chisholm RL. One stop shop for everything Dictyostelium: dictyBase and the Dicty Stock Center in 2012. Methods Mol Biol. 2013;983:59–92. doi: 10.1007/978-1-62703-302-2_4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Flicek P, et al. Ensembl 2014. Nucleic Acids Res. 2014;42(Database issue):D749–D755. doi: 10.1093/nar/gkt1196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Aurrecoechea C, et al. ApiDB: Integrated resources for the apicomplexan bioinformatics resource center. Nucleic Acids Res. 2007;35(Database issue):D427–D430. doi: 10.1093/nar/gkl880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Keeling PJ, et al. The Marine Microbial Eukaryote Transcriptome Sequencing Project (MMETSP): Illuminating the functional diversity of eukaryotic life in the oceans through transcriptome sequencing. PLoS Biol. 2014;12(6):e1001889. doi: 10.1371/journal.pbio.1001889. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Sun S, et al. Community cyberinfrastructure for Advanced Microbial Ecology Research and Analysis: The CAMERA resource. Nucleic Acids Res. 2011;39(Database issue):D546–D551. doi: 10.1093/nar/gkq1102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Emanuelsson O, Nielsen H, Brunak S, von Heijne G. Predicting subcellular localization of proteins based on their N-terminal amino acid sequence. J Mol Biol. 2000;300(4):1005–1016. doi: 10.1006/jmbi.2000.3903. [DOI] [PubMed] [Google Scholar]
- 55.Emanuelsson O, Brunak S, von Heijne G, Nielsen H. Locating proteins in the cell using TargetP, SignalP and related tools. Nat Protoc. 2007;2(4):953–971. doi: 10.1038/nprot.2007.131. [DOI] [PubMed] [Google Scholar]
- 56.Park JS, Simpson AG. 2011. Characterization of Pharyngomonas kirbyi (= “Macropharyngomonas halophila” nomen nudum), a very deep-branching, obligately halophilic heterolobosean flagellate. Protist 162(5):691–709.
- 57.Harding T, et al. Amoeba stages in the deepest branching heteroloboseans, including Pharyngomonas: Evolutionary and systematic implications. Protist. 2013;164(2):272–286. doi: 10.1016/j.protis.2012.08.002. [DOI] [PubMed] [Google Scholar]
- 58.Edgar RC. MUSCLE: A multiple sequence alignment method with reduced time and space complexity. BMC Bioinformatics. 2004;5:113. doi: 10.1186/1471-2105-5-113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Katoh K, Misawa K, Kuma K, Miyata T. MAFFT: A novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 2002;30(14):3059–3066. doi: 10.1093/nar/gkf436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Katoh K, Kuma K, Toh H, Miyata T. MAFFT version 5: Improvement in accuracy of multiple sequence alignment. Nucleic Acids Res. 2005;33(2):511–518. doi: 10.1093/nar/gki198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–780. doi: 10.1093/molbev/mst010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Criscuolo A, Gribaldo S. BMGE (Block Mapping and Gathering with Entropy): A new software for selection of phylogenetic informative regions from multiple sequence alignments. BMC Evol Biol. 2010;10:210. doi: 10.1186/1471-2148-10-210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Stamatakis A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30(9):1312–1313. doi: 10.1093/bioinformatics/btu033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Le SQ, Dang CC, Gascuel O. Modeling protein evolution with several amino acid replacement matrices depending on site rates. Mol Biol Evol. 2012;29(10):2921–2936. doi: 10.1093/molbev/mss112. [DOI] [PubMed] [Google Scholar]
- 65.Lartillot N, Lepage T, Blanquart S. PhyloBayes 3: A Bayesian software package for phylogenetic reconstruction and molecular dating. Bioinformatics. 2009;25(17):2286–2288. doi: 10.1093/bioinformatics/btp368. [DOI] [PubMed] [Google Scholar]
- 66.Shimodaira H, Hasegawa M. CONSEL: For assessing the confidence of phylogenetic tree selection. Bioinformatics. 2001;17(12):1246–1247. doi: 10.1093/bioinformatics/17.12.1246. [DOI] [PubMed] [Google Scholar]
- 67.Price DC, et al. Cyanophora paradoxa genome elucidates origin of photosynthesis in algae and plants. Science. 2012;335(6070):843–847. doi: 10.1126/science.1213561. [DOI] [PubMed] [Google Scholar]
- 68.Simpson AG, Patterson DJ. On core jakobids and excavate taxa: The ultrastructure of Jakoba incarcerata. J Eukaryot Microbiol. 2001;48(4):480–492. doi: 10.1111/j.1550-7408.2001.tb00183.x. [DOI] [PubMed] [Google Scholar]
- 69.Stairs CW, et al. A SUF Fe-S cluster biogenesis system in the mitochondrion-related organelles of the anaerobic protist Pygsuia. Curr Biol. 2014;24(11):1176–1186. doi: 10.1016/j.cub.2014.04.033. [DOI] [PubMed] [Google Scholar]
- 70.Miyagishima SY, et al. Two types of FtsZ proteins in mitochondria and red-lineage chloroplasts: The duplication of FtsZ is implicated in endosymbiosis. J Mol Evol. 2004;58(3):291–303. doi: 10.1007/s00239-003-2551-1. [DOI] [PubMed] [Google Scholar]
- 71.TerBush AD, Yoshida Y, Osteryoung KW. FtsZ in chloroplast division: Structure, function and evolution. Curr Opin Cell Biol. 2013;25(4):461–470. doi: 10.1016/j.ceb.2013.04.006. [DOI] [PubMed] [Google Scholar]
- 72.Heidel AJ, et al. Phylogeny-wide analysis of social amoeba genomes highlights ancient origins for complex intercellular communication. Genome Res. 2011;21(11):1882–1891. doi: 10.1101/gr.121137.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Romeralo M, Cavender JC, Landolt JC, Stephenson SL, Baldauf SL. An expanded phylogeny of social amoebas (Dictyostelia) shows increasing diversity and new morphological patterns. BMC Evol Biol. 2011;11:84. doi: 10.1186/1471-2148-11-84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Lara E, Chatzinotas A, Simpson AGB. Andalucia (n. gen.): The deepest branch within jakobids (Jakobida; Excavata), based on morphological and molecular study of a new flagellate from soil. J Eukaryot Microbiol. 2006;53(2):112–120. doi: 10.1111/j.1550-7408.2005.00081.x. [DOI] [PubMed] [Google Scholar]
- 75.Brown MW, Sharpe SC, Silberman JD, Heiss AA, Lang BF, et al. Phylogenomics demonstrates that breviate flagellates are related to opisthokonts and apusomonads. P Roy Soc B-Biol Sci. 2013;280(1769) doi: 10.1098/rspb.2013.1755. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Heiss AA, Walker G, Simpson AG. The microtubular cytoskeleton of the apusomonad Thecamonas, a sister lineage to the opisthokonts. Protist. 2013;164(5):598–621. doi: 10.1016/j.protis.2013.05.005. [DOI] [PubMed] [Google Scholar]
- 77.Barberà MJ, et al. Sawyeria marylandensis (Heterolobosea) has a hydrogenosome with novel metabolic properties. Eukaryot Cell. 2010;9(12):1913–1924. doi: 10.1128/EC.00122-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Bendich AJ, Gauriloff LP. Morphometric analysis of cucurbit mitochondria: The relationship between chondriome volume and DNA content. Protoplasma. 1984;119:1–7. [Google Scholar]
- 79.Seguí-Simarro JM, Coronado MJ, Staehelin LA. The mitochondrial cycle of Arabidopsis shoot apical meristem and leaf primordium meristematic cells is defined by a perinuclear tentaculate/cage-like mitochondrion. Plant Physiol. 2008;148(3):1380–1393. doi: 10.1104/pp.108.126953. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Buetow DE. 1989. The mitochondrion. The Biology of Euglena: Subcellular Biochemistry and Molecular Biology, ed Buetow DE (Academic, San Diego), Vol 4, pp 247–314.
- 81.Hogan MJ, Yoneda C, Feeney L, Zweigart P, Lewis A. Morphology and culture of Toxoplasma. Arch Ophthalmol. 1960;64:655–667. doi: 10.1001/archopht.1960.01840010657006. [DOI] [PubMed] [Google Scholar]
- 82.van Dooren GG, et al. Development of the endoplasmic reticulum, mitochondrion and apicoplast during the asexual life cycle of Plasmodium falciparum. Mol Microbiol. 2005;57(2):405–419. doi: 10.1111/j.1365-2958.2005.04699.x. [DOI] [PubMed] [Google Scholar]
- 83.Hardham AR. Ultrastructure and serial section reconstruction of zoospores of the fungus Phytophthora cinnamomi. Exp Mycol. 1987;11(4):297–306. [Google Scholar]
- 84.Yoshida Y, et al. The bacterial ZapA-like protein ZED is required for mitochondrial division. Curr Biol. 2009;19(17):1491–1497. doi: 10.1016/j.cub.2009.07.035. [DOI] [PubMed] [Google Scholar]
- 85.Friedman JR, et al. ER tubules mark sites of mitochondrial division. Science. 2011;334(6054):358–362. doi: 10.1126/science.1207385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Miyagishima SY, Nishida K, Kuroiwa T. An evolutionary puzzle: chloroplast and mitochondrial division rings. Trends Plant Sci. 2003;8(9):432–438. doi: 10.1016/S1360-1385(03)00193-6. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





