Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Population-Level Analyses Indirectly Reveal Cryptic Sex and

Life History Traits of Pseudoperkinsus tapetis (Ichthyosporea,


Opisthokonta): A Unicellular Relative of the Animals
Wyth L. Marshall*,1,2 and Mary L. Berbee1,2
1
Department of Botany, University of British Columbia, Vancouver, BC, Canada
2
Bamfield Marine Sciences Centre, Bamfield, BC, Canada
*Corresponding author: E-mail: WythMarshall@yahoo.ca.
Associate editor: Jeffrey Thorne

Abstract
Research article

We use population genetics to detect the molecular footprint of a sexual cycle, of a haploid vegetative state, and of lack of
host specificity in Pseudoperkinsus tapetis, a marine unicellular relative of the animals. Prior to this study, complete life
cycles were not known for any of the unicellular lineages sharing common ancestry with multicellular animals and fungi.
We established the first collection of conspecific cultures of any member from the unicellular opisthokont lineage
ichthyosporea, isolating 126 cultures of P. tapetis from guts of marine invertebrates ranging from clams to sea cucumbers.

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


We sequenced fragments of the elongation factor alpha-like (EFL) and heat-shock protein 70 (HSP70) genes for a subset of
our isolates. Absence of heterozygotes from the EFL locus in 52 isolates provided evidence for haploidy. Phylogenetic
incongruence and a lack of support for linkage between two loci from 34 sequenced isolates signified a history of
recombination consistent with a sexual cycle. Shared haplotypes in different invertebrate species showed that P. tapetis
was not host specific. Based on estimates of the frequency of sex and on observations of cultures, we propose that P.
tapetis is transmitted between hosts via asexual endospores. New protists are continually being discovered, and, as this
study illustrates, analysis of culturable collections from natural habitats can transform a species from a near unknown to
a model system for better understanding the evolution of life histories.
Key words: effective population size, ichthyosporea, mesomycetozoa, protist, Pseudoperkinsus, recombination.

Introduction cycles have yet to be elucidated (Maldonado 2005; Carr


Traditionally, determination of microbial life cycles has re- et al. 2009). Reasons for this include their recent discovery
lied on tenacity, microscopy, and real-time observation. and a lack of available cultures.
Our cultures of P. tapetis constitute one of the first col-
Using population genetic techniques and samples of con-
lections of conspecific isolates from any of the four new
specific isolates, we have worked in ‘‘reverse’’ describing as-
lineages of unicellular metazoan relatives. Previously, P.
pects of an organism’s life cycle and ecology based on the
tapetis was cultured only twice, once from Japan and once
footprints left in DNA sequence variation.
from Spain (Figueras et al. 2000; Takishita et al. 2008).
During a survey of marine invertebrate digestive tracts
Rather than adapt cytological methods to a divergent
for novel and culturable unicellular osmotrophs, we cul- and possibly host-adapted organism, we applied popula-
tured 126 isolates of an ichthyosporean with 99.9% se- tion-level genetic analyses to our collection of isolates.
quence similarity in the ribosomal small subunit rDNA Using a primer-based approach, we found two loci with
gene (SSU) to the shellfish symbiont Pseudoperkinsus tape- intraspecific variation, which we used to indirectly reveal
tis (Figueras et al. 2000; Takishita et al. 2008). The ichthyo- the ploidy, or chromosome copy number, of our isolates
sporea are one of seven unicellular lineages belonging to and detect genetic evidence of recombination. Using meas-
the eukaryote supergroup opisthokonta (Cavalier-Smith ures of nucleotide diversity, we estimated the effective pop-
1987; Adl et al. 2005). Prior to 1995, choanoflagellates were ulation size and frequency of recombination. We also
the only known unicellular relatives of animals (Cavalier- describe an enquiry into P. tapetis’s demographics and
Smith 1987). Since that time, four additional lineages have its host specificity.
been discovered (Adl et al. 2005). Of these new additions,
the ichthyosporea are, so far, the most speciose (Mendoza Genetic Evidence of Chromosome Copy Number
et al. 2002). Molecular phylogenetic analyses place the or- Ploidy is integral to a life cycle, but characterizing ploidy by
igin of the ichthyosporea at the basal-most node on the cytological observation can be difficult to adapt to new
animal lineage after the animal/fungus divide (Ruiz-Trillo groups of organisms. Flow cytometry could be used to sort
et al. 2006, 2008; Steenkamp et al. 2006; Shalchian-Tabrizi haploid from diploid cells (e.g., Bernander et al. 2001), except
et al. 2008). For the unicellular relatives of animals including that we had no means to establish baseline values for 1N and
choanoflagellates and ichthyosporea, ploidy and sexual 2N chromatin amounts. A widely adaptable approach to

© The Author 2010. Published by Oxford University Press on behalf of the Society for Molecular Biology and Evolution. All rights reserved. For permissions, please
e-mail: journals.permissions@oxfordjournals.org

2014 Mol. Biol. Evol. 27(9):2014–2026. 2010 doi:10.1093/molbev/msq078 Advance Access publication April 1, 2010
Cryptic Sex and Life History of Pseudoperkinsus · doi:10.1093/molbev/msq078 MBE
determining ploidy has been analysis of the frequency of het-
erozygotes in microsatellite data sampled from populations
(e.g., Oliveira et al. 1998; Brondani et al. 2000; Rynearson and
Armbrust 2000; van der Verde et al. 2001; Weeks et al. 2001).
The lack of heterozygotes from microsatellite data showed,
for example, that a population of Symbiodinium dinoflagel-
lates was haploid (Santos and Coffroth 2003). A similar prin-
ciple can be applied using sequence data from a population
with polymorphisms. Direct sequencing of polymerase chain
reaction (PCR) products from heterozygous diploids produ-
ces sequence chromatograms with double peaks at the seg-
regating sites (Clark 1990). The multicellular animals are
mainly diploid, and the higher fungi are mostly haploid or
dikaryotic. To assess chromosome copy numbers in P. tapetis,
we carefully screened the segregating sites of our sequence
chromatograms for double peaks, indicative of heterozy-
gotes and diploidy.
FIG. 1. Pseudoperkinsus tapetis isolation sites from Vancouver Island,

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


Genetic Evidence of Recombination in British Columbia, Canada. Western sites were distributed within
Barkley Sound, near the Bamfield Marine Sciences Centre. East side
Microeukaryotes sites were from three beaches along a 3 km SSW shoreline of
As in most microorganisms, asexuality is easily observed in Gabriola Island. Specific locations are listed in table S1.
the few ichthyosporeans that grow in culture (Whisler
1968; Jostensen et al. 2002; Marshall et al. 2008). Meiosis sociation between these loci, indicative of either a clonal
has yet to be reported in any ichthyosporean. Nuclear fu- population or close physical arrangement of the genes.
sion between coalesced uninucleate cells from adjacent fil-
aments, resembling the ladder-like conjugation of the green Life Cycle and Relationship with Host
alga Spirogyra, has been observed in the elongate ichthyo- Population-level analyses have been rare for the protozoa
sporean Enteropogon sexuale (Hibbits 1978). However, the and are lacking altogether for the unicellular opisthokont
fused cells did not grow or divide, and because E. sexuale is lineages. Within the ichthyosporea, no genetic studies have
host obligate, Hibbit’s observations were limited to short- been conducted to look further at population structure or
term study of dissected animals. Carr et al. (2008) argued host affinities. In order to use population-level genetic var-
that the presence of long terminal repeat retrotransposons iation to indirectly characterize the life cycle of P. tapetis,
in the genome of the choanoflagellate Monosiga brevicollis we sequenced two protein-coding loci that exhibited intra-
indicated a sexual history because if the species had been specific variation. By screening sequences for heterozygos-
purely asexual, the retrotransposon build up would have led ity, we were able to determine the chromosome copy
to mutational meltdown and extinction. However, empirical number. With ploidy established, we were able to test
evidence showing a sexual cycle has yet to be uncovered for for evidence of a sexual cycle and estimate effective pop-
any member of the unicellular relatives of animals. ulation size using measures of nucleotide diversity. Using
Sexual recombination, while widespread, can be difficult limited evidence of recombination, we were able to esti-
to observe and is unknown in many unicellular protists and mate the frequency of sex by calculating the population
fungi (e.g., Reynolds 1993); however, the inability to find it recombination rate based on the footprint left by homol-
does not rule out its occurrence. On the other hand, even ogous recombination in our sequences. In order to deter-
when mating can be induced in the laboratory, it may not mine whether sex was obligate or facultative, we compared
necessarily occur in nature. Molecular analysis of variable the average time between recombination events with host
loci in populations has the advantage of being able to show age. Ultimately, we combined these findings with culture
that sexual reproduction has occurred in nature by reveal- and isolation observations and proposed a life cycle that
ing patterns originating from cryptic recombination within included the host/symbiont relationship.
natural populations of pathogenic and free-living bacteria,
fungi, and protozoa (Maynard Smith et al. 1993; Burt et al. Materials and Methods
1996; Geiser et al. 1998; LoBuglio and Taylor 2002; Xu 2004;
Campbell and Carter 2006; Cooper et al. 2007). Genomes of Animal Collection, Isolation of Strains and Culture
asexual organisms are inherited as effectively identical cop- Conditions
ies of the parent. Sexual genomes are shuffled and reas- We cultured P. tapetis from eight marine invertebrate host
sembled by independent assortment of chromosomes collections plus one sample of water surrounding captive
and crossing over between paired chromosomes. We tested oysters from the east and west sides of central Vancouver
whether P. tapetis had an incongruent phylogenetic history Island, British Columbia, Canada, between August 2004 and
between two loci, consistent with recombination. To the April 2007 (fig. 1 and table 1). Pseudoperkinsus tapetis was
opposite effect, we also tested for significant linkage, or as- not found from attempted isolations from an additional

2015
Marshall and Berbee · doi:10.1093/molbev/msq078 MBE
Table 1. Collection Location, Animal Host and Haplotype for the tion of a 150 ml aliquot of 0.5 l of water surrounding three
34 Sequenced Pseudoperkinsus tapetis Isolates. captive oysters. Ichthyosporean colonies were white and
Animal Host Haplotype raised and consisted of large round cells. We repeatedly
Location (number collected) Numbera Isolate IDb streaked each of the individual colonies from the starting
East 1 Mussel (5) 1* GM8 plates onto fresh media until they were axenic. For isolates
East 2 Mussel (5)c 1* MB4 CG2 and LE7, we observed asexual reproduction and stained
3 BM6, MB5, MB18
4 BM33
the nuclei with 4#,6-diamidino-2-phenylindole (DAPI) ac-
5* BM19 cording to Marshall et al. (2008). These strains are available
6 MB29 at the American Type Culture Collection (ATCC) under
12* MMB5, MMB3, MB13 accession numbers PRA-281/282.
13* BM17
14 MB16
Varnish clam (5)c 5* NO32
DNA Extraction, PCR Amplification, and
10 ON1M Sequencing
12* NO6, ON8, ON50M We extracted DNA from approximately 50–200 ll of cells
13* ON12 scraped from plates with confluent growth using a DNeasy
16 NO16 Plant Mini Kit (QIAGEN Inc., Mississauga, Ontario, Canada).
East 3 Manilla clam (5)d 2 PV3, PV38, VP2
3* PV25, VP7
We amplified and sequenced the SSU, internal transcribed
5* PV23 spacers ITS1 and ITS2 and the 5.8S ribosomal subunit rDNA

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


11 VP19, VP17 gene (ITS), and a region of the elongation factor alpha-like
West 1 Pacific oyster (3)c 7 CG2 gene (EFL) directly from PCR products (Marshall et al. 2008),
8 CG5 except that we used internal primers ELF3 and ELR3 (sup-
15 CRG9
West 2 Chiton (3) 9 KS1
plementary table S2, Supplementary Material online) to se-
West 3 Sea cucumber (5) 17 LE7 quence the EFL. For initial primer design, we amplified
West 4 Oyster water 18 OI22B a region of the cytosolic heat-shock protein 70 gene
NOTE.—East 1–3 refers to each of three beaches on Gabriola Island, and west 1–4 (HSP70) using primers H7150F and H7965R (supplementary
corresponds to four sites within Barkley Sound. Specific locations are provided in table S2, Supplementary Material online) with an annealing
table S1.
a temperature of 48 °C and then cloned the fragment into the
Asterisks mark haplotypes that were found in more than one host collection.
b
A representative of each haplotype per host collection was included in the vector pCR2.1 using the TOPO TA cloning kit (Invitrogen)
‘‘individual’’ data set (see figs. 2 and 3). and sequenced clones. We then amplified the HSP70 from
c
Animal collections where the number of haplotypes is equal to or greater than our isolates using our new internal primers H7F2 and H7R2
the number of animals pooled.
d
The number of haplotypes is at least 5 (see table S1). (supplementary table S2, Supplementary Material online)
with an annealing temperature of 52 °C and sequenced
the PCR products directly. Sequences were determined
32 animal collections (supplementary table S1, Supplemen- by Macrogen at Seoul, Korea, or the Nucleic Acid Protein
tary Material online). Because the geographic extent of Service Unit at the University of British Columbia, Canada.
populations of ichthyosporeans is unknown and incorrect To verify point mutations, we sequenced each allele from at
estimations of population size can sometimes give false in- least two PCR products and in both directions (supplemen-
dications of clonality (e.g., Campbell et al. 2005), we sam- tary table S1, Supplementary Material online), mixing all
pled animals from overlapping (cohabiting) sites as well as PCR reagents in a laminar flow hood to reduce or eliminate
beaches separated by 1–200 km of coastline (fig. 1, table 1, cross contamination. As a check, we reamplified and rese-
and supplementary table S1, Supplementary Material on- quenced 36% of HSP70 and 50% of EFL genes (fig. 2 and sup-
line). Animals from the east side were collected from three plementary table S1, Supplementary Material online) and
beaches, each separated by 1 km along the southern shore found no evidence for cross contamination. The haplotypes
of Gabriola Island. Animals from the west side were col- represented by isolates GM8, BM19, VP19, PV23, LE7, OI22B,
lected from four sites within Barkley Sound, each separated BM6, PV25, NO6, MMB5, ON1M, PV3, BM17, MB29, and
between 0.75 and 2.5 km. Specific locations are provided in ON12 were verified either through being reamplified and
supplementary table S1, Supplementary Material online. To resequenced or by being found in other isolates (supple-
obtain culture material, we pooled portions of the digestive mentary table S1, Supplementary Material online).
tracts from three to five animals from each of the eight host We worked under the assumption that conspecific iso-
collections (Marshall et al. 2008). For most animals, we used lates could be recognized and identified as P. tapetis based
pieces from all gut regions; for bivalves, we used a portion on a phylogenetic species concept. All P. tapetis isolates
of the digestive gland and syringed stomach material. We formed a monophyletic group. A subset of eight of our iso-
diluted the gut contents in 400 ll of sterile seawater and lates (supplementary table S1, Supplementary Material on-
inoculated culture plates with 10 and 30 ll of the slurry line) were 100% identical in SSU sequence to each other
(Marshall et al. 2008). As a control, we cultured undiluted over a maximum overlap of 1718 bp and were 99.9%
coelomic fluid (non-bivalve animals) and mantle cavity fluid identical to the only two other known P. tapetis isolates
(bivalves) from five of eight pooled animal collections. We (Figueras et al. 2000; Takishita et al. 2008). All 126 of
also isolated cultures from a pellet formed from centrifuga- our isolates shared 99–100% ITS identity to each other

2016
Cryptic Sex and Life History of Pseudoperkinsus · doi:10.1093/molbev/msq078 MBE
and were excluded from this study because they were at
most 88.4% identical to P. tapetis over a 725-bp region of
overlap in their ITS regions. These levels of identity were
comparable with the 5.1% interspecific versus 0.51% intra-
specific ITS divergence within a data set of Penicillium
species (Seifert and Samson 2007). Of our 126 putatively
conspecific isolates, we chose 34 isolates for sequencing
of the EFL and HSP70 (table 1). These isolates represented
a selection from all eight animal hosts (table 1 and supple-
mentary table S1, Supplementary Material online). Because
inbreeding can result in elevated levels of homozygosity, we
increased the number of sequences from the EFL locus to 52.

Genetic Analysis
We aligned sequences in ClustalX v1.8 (Thompson et al.
1997) and edited them in MacClade 4.08 (Maddison WP
and Maddison DR 2000). We calculated pairwise nucleotide
diversity, p, and theta per site, hw (Watterson 1975), in

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


DnaSP v.4.50.3 (Rozas et al. 2003) based on a data set of
24 individuals (defined below). To construct parsimony
trees, we used TCS v.1.13 (Clement et al. 2000) and PAUP*
v.4.0b10 (Swofford 1998). Parsimony bootstrapping in
PAUP* involved 500 heuristic search replicates with 10 ran-
dom sequence additions with tree-bisection reconnection
(TBR) branch swapping limited to 106 rearrangements. We
used the most parsimonious tree from each locus to cal-
culate the consistency index (CI), retention index (RI), and
FIG. 2. The genealogical network from the HSP70 locus (top) tree length using PAUP*.
conflicts with the network from the EFL locus (bottom), suggesting To decide chromosome copy number, we checked orig-
a history of recombination between the two loci. Each alphanu- inal chromatograms for each of the EFL sequences, looking
meric code designates an individual isolate; isolates collected from for double peaks at any of the 19 segregating sites identified
the west side are identified by italics (for isolate details, see fig. 3 and
table 1). Code color specifies membership in one of three HSP70
after aligning all 52 EFL sequences.
haplotype networks, designated in blue, black, and red. Instead of
the expected pattern in the absence of recombination, where Definition of Data Sets
isolates of the same color would generally group together in both In order to eliminate clone siblings that may have been iso-
the HSP70 and the EFL networks, the rearrangement of the color lated during dissection and culturing, we included only
patterns in the EFL network suggests recombination. Networks were unique haplotypes from each animal collection (hereafter
generated by TCS (Clement et al. 2000). Numbers above branches referred to as ‘‘individuals’’) in our first data set (n 5 24).
are branch lengths from TCS; numbers below are parsimony
bootstrap percentages from PAUP* (Swofford 1998). Consistency
Asexual reproduction was obvious in culture. For analyses
index (C.I.), Retention Index (R.I.), segregating sites (S), and tree where a history of sexual recombination could be obscured
length (L) are shown for both trees. Dashed lines between the three by including asexual siblings, we defined two additional
HSP70 networks show hypothetical connectivity after the TCS data sets with varying degrees of correction to eliminate
connection limit was increased to 60 steps. This diagram also possible clones. Some isolates came from animals collected
illustrates three alternative approaches to designating alleles for from the same beach. Assuming clones dispersed locally, we
further analysis. Circles around isolates with identical sequences used unique haplotypes from the each location for a second
designate 11 HSP70 and 10 EFL alleles. To minimize the effects of
recent somatic mutations, three alternative approaches to desig-
‘‘location-corrected’’ data set (n 5 21). In order to assess
nating alleles were used: 1) All sites were considered (circles 1–10 or the relationship of the loci independent of population
11), 2) sequences with only one unique difference were lumped structure, we used a third data set consisting of unique
together as the same allele (A–F or H, dashed boxes), or 3) alleles haplotypes (n 5 18). In addition to allowing for alternative
were defined using only balanced sites (fig. 3) (solid boxes E/H1–5). degrees of stringency in removing possible clones, we also
applied three alternative definitions of alleles for analyses
and to the single isolate of P. tapetis from Spain over a re- where results depended on how stringently the alleles are
gion with a maximum overlap of 667 bp including ambig- defined (see fig. 2).
uous sites with secondary peaks. Compared with the P.
tapetis isolate from Japan, our isolates shared 96.9–98.8% Expected Number of Individuals
ITS similarity over a maximum overlap of 675 bp. In con- Animals collected from the east side of Vancouver Island
trast, other ichthyosporeans that we isolated (data not gave rise to a high number of isolates. When isolates with
shown) were not considered to be conspecific with P. tapetis identical haplotypes were cultured from the same animal

2017
Marshall and Berbee · doi:10.1093/molbev/msq078 MBE

FIG. 3. Polymorphic sites from the amplified regions of the EFL and HSP70 loci shown for each of 24 individuals or unique haplotypes per host
collection. Each individual was assigned 1 of 18 unique haplotypes defined after the EFL and HSP70 sequences were concatenated. Horizontal lines
separate isolates based on their combined haplotype designation. A small ‘‘g’’ denotes a gap, and each gap was coded as one character regardless of
its length. The noncoding region of the HSP70 is shaded in grey. The polymorphic region used to calculate the expected heterozygosity of the EFL is
outlined by a box. Balanced sites were defined as any site where polymorphisms were present in at least three individuals and are marked by an
asterisk above the consensus line. Allele and haplotype numbers correspond to figure 2. Specific collection locations are listed in table S1.

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


collection, it was impossible to know whether they origi- Burt 2001). Values range between 0 for randomly associ-
nated from one or more initial infections of the host. How- ated alleles between loci (due to panmictic mating) and
ever, assuming that haplotypes infected randomly and that 1 for complete linkage. To assess significance of the alter-
infections were Poisson distributed based on average hap- nate hypothesis of clonality (P , 0.05), we compared the
lotype frequency, we used an arithmetic approach to esti- observed data set with 10,000 replications that mimicked
mate a fourth data set, the actual or ‘‘expected’’ number of the effects of random mating. The IA was calculated and
infective individuals based on our sequenced ‘‘individual’’ tested for all three allele definitions (fig. 2) for the individ-
isolates from the east side. Where f was the number of hosts ual, location corrected, and unique haplotype data sets (ta-
infected with the same haplotype, f/4 was the observed fre- ble 2). We also calculated and tested the IA for the
quency of finding the haplotype among our four host col- individuals and unique haplotypes on the east side (table
lections, ln (1  f/4) was the Poisson-corrected expected 2).
frequency of the haplotype infection of four host collec- The parsimony tree length permutation test (PTLPT)
tions, and finally (ln (1  f/4)*4) was the number of times (Archie 1989; Burt et al. 1996) uses all the sequence data
that the same haplotype of P. tapetis would have been ex- and therefore takes into account similarity among alleles
pected to independently infect the pool of four hosts. If, for as well as allele frequency. For the PTLPT in the absence
example, the same haplotype of P. tapetis was isolated from of recombination, the combined tree length for both loci
two of the four hosts, its observed frequency was 0.5, its should be significantly shorter (P , 0.05) than trees gener-
expected frequency was 0.69, and the expected number ated after simulating random mating with random mixing of
of cells independently initiating infection was 2.77. the alleles. We generated 1,000 randomizations using Multi-
locus with alleles coded as character states for each locus for
Tests for Evidence of Clonality and Recombination the individual and unique haplotype data sets. Following
randomization, we replaced the allele codes with the orig-
Genetic Tests for Clonality
inal sequence data and calculated the tree lengths in PAUP*
Linkage disequilibrium or nonrandom association among
according to the specifications provided by Multilocus.
alleles between loci can arise from genetic isolation, in-
breeding or lack of recombination, and clonal expansion. Tests for Recombination
We tested the null hypothesis of recombination by calcu- Sexual recombination in eukaryotes is a consequence of
lating the index of association (IA) (Brown et al. 1980; May- segregation and independent assortment of chromosomes
nard Smith et al. 1993) using Multilocus v.1.3 (Agapow and during meiosis resulting in different genealogies for loci on

Table 2. Significant (P , 0.05) Indexes of Association (IA) Between HSP70 and EFL Alleles Were Evident When All Individual Isolates Were
Included, But Not When Data Sets Consisted of Unique Haplotypes or Location Corrected Haplotypes.
Allele Definitions Isolates Included Number of Isolates IA P
All sites considered All individuals 24 0.1814 0.015
Alleles reduced by one variable site (fig. 2) All individuals 24 0.1482 0.025
Alleles defined by balanced sites (fig. 2) All individuals 24 0.1328 0.036
All sites considered Location corrected 21 0.0717 0.23
All sites considered Unique haplotypes 18 20.0736 1.0
All sites considered East side individuals 18 0.1949 0.034
All sites considered East side unique haplotypes 12 20.1277 1.0

2018
Cryptic Sex and Life History of Pseudoperkinsus · doi:10.1093/molbev/msq078 MBE
separate chromosomes. We tested for incongruent histories The population recombination parameter (q) is the mea-
between the HSP70 and the EFL using the incongruence length sure of overall covariation of nucleotides at different sites,
difference (ILD) test (Farris et al. 1994) for the data set of 24 and (rr) is a compound value based on the observed num-
individuals. The ILD tests a null hypothesis of clonality by com- ber of chiasmata in Morgans per kilobase (M/kb) during
paring the summed tree length of the observed data with the meiotic division or per meiosis recombination rate (r) mul-
sums after random partitioning. We generated 1,000 trees in tiplied by the frequency of sex (r). We used the HSP70 intron
PAUP* using heuristic searches with random sequence addi- to calculate hw to avoid the effects of selection and codon use
tion and TBR branch swapping limited to 106 rearrangements. bias. We estimated the per meiosis recombination rate (r) as
Our cut-off for significance was P , 0.01, and we excluded in- being between 0.000017 and 0.0048 M/kb based on obser-
variant characters (Cunningham 1997). vations from unicellular eukaryotes (Awadella 2003; Lynch
To map the history of recombination between the 2006). An r value of 103 M/kb corresponds to roughly 1
two loci, we made recombination networks (Huson and chiasma for a 1 Mb chromosome. We did not account for
Kloepper 2005) for the EFL, HSP70, and both regions com- inbreeding nor consider the series of mitotic divisions be-
bined in Splits Tree4 v4.10 using the data set of 24 individ- tween endospore growth and release of new daughter cells.
uals (Huson 1998; Huson and Bryant 2006). To minimize Because we had so few isolates from our west side sites, we
homoplasies and keep the number of segregating sites were unable to confidently test for population connectivity.
within each locus comparable, we only used sequence from However, alleles from both loci were shared by isolates from
the second exon of HSP70 (fig. 3) when the loci were com- both sides of the island, and we assumed that our 24 indi-

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


bined. We tested for recombination within each locus and viduals were part of a single population.
within the combined loci using the ‘‘pairwise homoplasy in-
dex’’ (PHI) (Bruen et al. 2006) implemented in Splits Tree4. Results
Homologous recombination is a second consequence of
meiosis and measurements of covariation between nucleo- ITS Variation
tides provides a basis for estimating the frequency of sexual Of a total of 975 sequenced sites within the ITS, 29 were
recombination. We tested both loci using the PHI test frequently ambiguous, with secondary peaks clearly visible
(as above). We also compared alternate methods, likeli- within cleanly sequenced regions. However, resolution and
hood analysis of recombination in DNA (LARD) (Holmes strength of the second peaks at the polymorphic sites were
et al. 1999), and Bootscan (Martin, Posada, et al. 2005), im- inconsistent between sequence reads from single isolates.
plemented by RDP3 software (Martin, Williamson, et al. In most eukaryotes, ribosomal regions including the ITS are
2005) after excluding alleles with single unique mutations. present in multiple tandem repeats. The patterns of se-
To measure the degree of covariation of nucleotides at dif- quence polymorphism in the ITS were consistent with var-
ferent sites within our 24 individuals, we calculated the iation among repeats rather than allelic variation, and the
population recombination parameter (q) from the inter- ITS variants were not analyzed further. When we ignored
genic region of the HSP70 locus using the moment method ambiguous sites, 123 isolates (including CG2) were identi-
of Hudson (1987) and R minimum of Hudson and Kaplan cal and 3 others, including LE7 and OI22B, differed at three
(1985) in DnaSP. We also used LDhat v.2.1 (McVean et al. sites. A single locus such as the ITS is not sufficient to define
2002) to calculate q using Wakeley’s moment method a species, and our population could still have included hy-
(Wakeley 1997) and the pairwise approximate likelihood brids or more than one species. Isolates LE7 and OI22B were
method (McVean et al. 2002). the most divergent in their ITS, EFL, and HSP70 sequences.
We repeated tests for index of association, PTLPT, and ILD
Effective Population Size
without LE7 and OI22B, but the significance of our results
We calculated the effective population size (Ne) according
was unchanged.
to Ne 5 hw/2l. The upper and lower bounds of the mu-
tation rate per base pair per mitosis (l) were 0.22–2.5 
109 based on ranges from unicellular eukaryotes (Lynch
Population Variation and Lack of Heterozygosity
The EFL data matrix, based on the 24 individual isolates,
2006). We calculated overall per site nucleotide variation
had 922 nucleotides with 19 segregating sites and 10 alleles
(hw) from the HSP70 intron and the 4-fold degenerate sites
(figs. 2 and 3). The amino acid translation was 307 codons
from the coding regions of HSP70 and EFL for all 24 indi-
in length, and all but one nonsynonymous substitution
viduals. For comparison, we also calculated hw and Ne for
were in the first or second position. All synonymous sub-
the EFL using the east and west side isolates separately.
stitutions were in the third position. The HSP70 matrix,
Frequency of Sex based on 24 individuals, consisted of 1,006 nucleotides with
We estimated the frequency of sex by comparing the rel- 64 segregating sites and 11 alleles (figs. 2 and 3). Two coding
ative contributions of same site variation due to mutation regions with a total of 13 synonymous substitutions flanked
and covariation of nucleotides due to recombination to a 524 bp intron (fig. 3).
overall nucleotide diversity (Tsai et al. 2008). In the stan- If our isolates were diploid, double peaks indicative of
dard neutral model, Ne can be calculated from the popu- heterozygotes should have been evident as conflicting sig-
lation mutation parameters (as above) or recombination nals on DNA sequence chromatograms. For example, as-
parameters according to Ne 5 q/2(rr) (Tsai et al. 2008). suming Hardy–Weinberg equilibrium, the frequency of

2019
Marshall and Berbee · doi:10.1093/molbev/msq078 MBE
heterozygotes at a chosen site within the EFL (fig. 3) should
have been 56.9% based on the frequencies of the alleles in
the individual data set. However, the observed frequency of
heterozygotes among the 52 isolates sequenced for the EFL
was zero. There were also no double peaks at any of the
polymorphic sites within the 34 isolates sequenced for
the HSP70. We therefore conducted all further analyses un-
der the premise that P. tapetis was haploid.
We found six unique haplotypes, one time each, from the
west side of Vancouver Island, and all the other isolates came
from four host collections from three beaches within a 2 km
range on the east side of the island (table 1 and fig. 1). The 12
different haplotypes found on the east side represented
a minimum of 18 independent host colonizations. The num-
ber was a minimum because it was calculated by ignoring
repeated cultures with the same haplotype from the same
host collection under the assumption that genetically iden-
tical cells originated and spread from a single initial coloni-

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


zation. However, in some instances, cells of the same
haplotype may have colonized independently. Factoring
in the probability of multiple independent infections by
the same haplotype increased the possible number of inde-
pendent colonizations in the east side hosts from 18 to 25.
Haplotypes were not correlated with host species. In-
stead, the same haplotypes were found in more than
one host species (table 1). An excess of P. tapetis haplotypes
relative to number of host animals pooled for culture in-
dicated that some host animals contained more than one
haplotype (table 1). One host sample of five animals had
eight different P. tapetis haplotypes, and several animal col-
lections had a number of haplotypes equal to the number
of animals sampled (table 1).

Evidence of Asexual Reproduction and Clonality


Morphological and Genetic Evidence of Clonality
Uninucleate endospores grew into multinucleate plasmodia
before maturation (fig. 4a). DAPI-stained cells undergoing
asexual reproduction in culture typically had 32, 64, or more FIG. 4. Asexual reproduction in Pseudoperkinsus tapetis. Like most
nuclei by the time endospores were visible. Numbers higher ichthyosporeans, P. tapetis produces a multinucleate parent cell,
than 64 nuclei per cell could not be accurately counted. The followed by the release of asexually derived endospores. (a) (i–vi)
rare cells that had nuclear counts on the order of 250 were DAPI stain of nuclear proliferation in P. tapetis. Cells start as
often irregular in shape and did not appear to have endo- uninucleate endospores, nuclei divide synchronously producing 2, 4,
8, 16, and so forth, multinucleate stages. Scale bar equals 20 microns.
spores. Nuclear division was synchronous, and nuclei in ma-
(b and c) Uninucleate nonmotile endospores are released (b) when
ture cells were the product of five to nine mitotic divisions. the parent cell wall splits into two valves or (c) by squeezing through
After 3–5 days, endospores were usually released through pores in the parent cell wall. Scale bars equal 20 and 10 microns.
pores or splitting of the parent cell wall (fig. 4b).
The population-level footprint of clonal or asexual re-
production was evident as a significant IA value when all
mutated data set including all individuals; however, the null
individuals were included (table 2). The IA decreased as al-
hypothesis of recombination could not be rejected using
lele definitions were adjusted to encompass minor variants
the data set consisting of unique haplotypes (P . 0.078).
but still indicated significant clonality (table 2). However,
the null hypothesis of recombination could not be rejected
when the sample was reduced to the location-corrected Recombination, Effective Population Size, and
data set (table 2) or the third data set that consisted of Frequency of Sex
unique haplotypes for both sides or from the east side Genetic Evidence for Sex
alone (table 2). Similarly, when sequences were used in Interlocus phylogenetic conflict pointed to a history of re-
place of allele codes (PTLPT), the observed tree length combination in the population. The TCS parsimony net-
was significantly shorter (P , 0.023) compared with a per- works from each locus (fig. 2) were incongruent, and

2020
Cryptic Sex and Life History of Pseudoperkinsus · doi:10.1093/molbev/msq078 MBE

FIG. 5. Inter- and intralocus recombination reconstructed using Splits Tree4. Reticulations representing potential recombination events are

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


shown in bold grey. Networks are not drawn to scale. (a) Networks from the combined EFL plus the HSP70 exon 2 region, compared with the
exon 2 region of HSP70 region alone (inset). The dashed lines represents reticulations shared by both trees. (b) Recombination network using
all 1,006 base pairs from our HSP70 sequence.

the tree lengths of the combined data sets were not addi- the east side and the intergenic region of the HSP70 had
tive. The C.I., R.I., and lengths calculated in PAUP* for each higher diversity than the degenerate sites from the coding
locus were 0.94, 0.99, and 68 for the HSP70 and 1.0, 1.0, and regions (table 4). Using intergenic and degenerate sites, the
20 for the EFL, respectively. The values for the combined Ne was on the order of 106 to 107 depending on the mutation
data sets were C.I. 5 0.8077, R.I. 5 0.9373, and length 5 rate (table 4).
104. When the two loci were analyzed separately, the ob- Frequency of Sex
served number of steps exceeded the minimum number of Ratios of sexual to asexual generations or frequency of sex,
steps only by 1 (EFL) or 4 (HSP70), but when the loci were r, ranged between 1 sexual cycle for 154–495,000 asexual or
combined, the tree was 21 steps longer than the minimum mitotic cycles with a median value of approximately once
and the ILD test was significant (P 5 0.001). Similarly, the for every 22,700 asexual cycles when q equalled 1 (fig. 6).
combined network for the coding region of the HSP70 and Median values were based on the two less extreme esti-
the EFL had several reticulations (fig. 5a), suggestive of re- mates. Our estimate of q was based on a relatively short
combination. A reticulation was detected when a fragment fragment of DNA, and q can vary dramatically between loci
of the second exon of the HSP70 was analyzed alone (fig. (Awadella 2003) for reasons including chromosomal posi-
5a), although without significant support from the PHI test tion. When we repeated our calculations using a hypothet-
(P 5 0.2), whereas the EFL had no reticulations. The PHI ical high q of 50, the frequency of sex ranged between 1 in 3
test for the combined data was significant (P , 0.01). and 1 in 9,910 asexual generations, with a median value of 1
Reticulations and intralocus recombination were evi-
dent in the entire sequenced region of the HSP70 (PHI:
P , 0.01) (fig. 5b). Both Bootscan and LARD detected three
significant recombination events (P  0.05) but neither Table 3. Estimates of Effective Population Size Times the
could confidently distinguish daughter sequences from Frequency of Sex (Ner)a Based on Several Estimates of Population
the minor parent or identify breakpoints. Both methods Recombination Parameter qb Calculated for the Data Set
suggested OI22B as a major parent, BM6 as a minor parent, Including All 24 Individuals.
and NO6 and CRG9 were two potential daughters. A third Effective Population Size 3
scenario was BM6 as a major parent, OI22B as a minor Frequency of Sex (Nes)a
parent, and LE7 a daughter. Method rb r/uw rb 5 0.000017 rb 5 0.0048
Effective Population Size Using hw Wakeley (1997) 0.285 0.0136 8,380 30
Rm/pairwise 1 0.044 29,400 104
Estimates of q for the HSP70 ranged between 0.285 and 1.6 Hudson (1987) 1.6 0.070 47,100 167
(table 3). Estimates of nucleotide diversity (p) and hw were
NOTE.—The q/hwc ratio was calculated using hw from the HSP70 intron and Ner
higher for the HSP70 than the EFL, although this difference based the upper and lower bounds of per meiosis recombination rate, rb.
was minimized when only 4-fold degenerate sites were con- a
(Ner) is the effective population size, Ne, times the frequency of sex, r. (Ner) is
sidered (data not shown). hw ranged between 0.0126 and calculated based on Ner 5 q/2r.
b
q and r are expressed in Morgans per kilobase (M/kb).
0.0227 using degenerate or intergenic sites (table 4). For c
hw is Watterson’s (1975) estimator of nucleotide diversity expressed per
the EFL, nucleotide diversity was higher on the west side than kilobase.

2021
Marshall and Berbee · doi:10.1093/molbev/msq078 MBE
Table 4. Effective Population Sizes (Ne) Based on Intergenic and 4-fold Degenerate Sites Given an Upper and Lower Bound for Mutation Rate l.
Locus and Sites Used (number of nucleotides) uw Ne Given m 5 0.22 3 1029 Ne Given m 5 2.5 3 1029
HSP70 coding (74) 0.0179 40,680,000 3,580,000
HSP70 intron (524) 0.0227 51,590,000 4,540,000
EFL coding, both sides (155) 0.0126 29,090,000 2,560,000
EFL coding, east side Vancouver Island (155) 0.0110 25,000,000 2,200,000
EFL coding, west side Vancouver Island (155) 0.0176 31,360,000 2,760,000
NOTE.—Nucleotide diversity (hw) was calculated from all 24 individuals.

sexual generation for 418 asexual generations. A data set tapetis that could not be observed in the field or in culture.
based on the expected number of infective individuals The complete absence of heterozygotes at polymorphic
as estimated with a Poisson correction for repeated infec- sites suggested that our P. tapetis population lives in its in-
tions by the same haplotype increased the number of in- vertebrate host primarily as a haploid and that it has a hap-
dividuals by 7 to 31. Because the expected data set included loid asexual cycle. If life cycles in other ichthyosporea are
a higher proportion of identical isolates, the resulting esti- predominantly haploid, Hibbits’ (1978) observations of
mates of the frequency of sex were slightly lower, with a me- cells of E. sexuale undergoing nuclear fusion could be inter-
dian value near 1 sexual cycle for 23,900 asexual cycles preted as fusing haploid cells in the process of forming a zy-

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


when q equalled 1. gote. No diploid asexual cycle was evident in P. tapetis,
which may indicate that it has a short-lived zygote, or
a patchy distribution of haploid and diploid populations,
Discussion or that diploids for some reason do not colonize animals
The combined results from microscopy and genetic anal- or grow in culture.
ysis from our collection of conspecifics enabled us to indi- An alternate explanation for a lack of observed hetero-
rectly unveil aspects of the life cycle and ecology of P. zygotes could be a diploid population that is highly inbred.
Johnson et al. (2003) found only one heterozygote among
28 isolates from a recombining population of the diploid
terrestrial yeast Saccharomyces paradoxus. For P. tapetis,
haploidy seems more likely than absolute inbreeding. Con-
sidering the large number of alleles, outcrossing would
have to be extremely rare to maintain 100% homozygosity.
No forms of cellular fusion such as those seen in yeasts
(Johannsen and van der Walt 1980; Johnson et al. 2003) were
observed under our culture conditions. Compared with
nonmotile terrestrial microorganisms, P. tapetis should have
more opportunities to disperse because cells are carried in
currents and tides and to outcross because host animals in-
gest cells and concentrate them. Based either on the min-
imum number of colonizations or on the calculated
expected number of colonizations, the number of infections
was greater than the number of animals pooled in several
samples. Because P. tapetis was not genetically isolated
within its host, it should have potential partners for
FIG. 6. Range of possible frequencies of sexual reproduction in outcrossing.
Pseudoperkinsus tapetis, given three estimates for the population
recombination factor (q) calculated for the heat-shock protein 70 Evidence for Both Clonality and Recombination
locus (x axis) where a y axis value of 1 is equivalent to no asexual Tests for clonality, such as the index of association, use pan-
reproduction. The vertical ranges represent alternative assumptions
about the per replication recombination rate (r) and mutation rate
mictic mating as a null hypothesis and showed that pan-
(l). Black symbols represent l 5 2.5  109 and white symbols mixis alone was not an adequate explanation for genetic
l 5 2.2  1010, and rectangles represent r 5 4.8  103 and structure. Additional evidence in favor of clonality was that
circles r 5 1.7  105. Median values (*) were calculated between all four gametic types were never found for any pair of al-
the two least extreme mutation and recombination rate combina- leles (Tibayrenc et al. 1990), but the absence of some of the
tions. If sexual reproduction was less frequent than once in 700 expected recombinant combinations may have been due
nuclear divisions (shaded region), and assuming that P. tapetis is to insufficient sampling. The lines of evidence for sexuality
retained in its host for a year, then initial host infection is probably
via asexual spores. The dotted lines are cut-offs, below which asexual
relied on the premise that under a clonal model, all regions
spores would constitute the infective stage given retention times or of the genome have the same history. Once our data were
host lifespans of 1, 5, and 10 years. Ninety percent inbreeding would reduced to unique haplotypes, the history of meiotic rear-
increase the range of frequency of sex by a factor of ;5. rangement was evident, and from the index of association,

2022
Cryptic Sex and Life History of Pseudoperkinsus · doi:10.1093/molbev/msq078 MBE
we were unable to reject a null hypothesis of recombina- rounds of mitosis per sporangium, the ratio of frequency
tion. The EFL and HSP70 had significantly incongruent his- of meiotic divisions to asexual endospore-to-endospore
tories, an expected consequence of a meiotic stage with generations, would be six times the meiosis:mitosis ratio.
independent chromosome assortment (or recombination For example, in figure 6, with q 5 1.0, the lowest estimate
between loci on the same chromosome). Evidence of intra- of meiosis:mitosis was 1:154, but the ratio of meiosis:asex-
locus recombination in the HSP70 provided more positive ual endospore-to-endospore generations would be 1:25.7.
support for a sexual cycle in our haploid organism where This last ratio would imply that P. tapetis undergoes at least
meiosis would have offered the only opportunity for cross- 25 rounds of sporangium formation for each round of mei-
over between different alleles from a single copy homolo- osis. The change in the units of measure for an asexual
gous gene. We did not find intralocus recombination in the generation would, of course, have no effect on the time
EFL, but this was unsurprising as intralocus recombination between sexual generations or the probability of infection
is less common than interlocus recombination and the of a host by an asexual propagule (fig. 6).
EFL was less variable than the HSP70 (fig. 3), so the prob- Regardless of the margins of error, P. tapetis is most likely
ability of detecting recombination within the EFL was facultatively sexual and can probably undergo unlimited
lower. Although difficult to observe directly, the evidence rounds of asexual reproduction, much like a yeast. Out-
for recombination was not unexpected. Meiosis is likely an crossing rates in yeasts have been estimated at once per
integral process in all extant eukaryotes except for some 3,000–9,000 asexual cycles for S. paradoxus (Tsai et al.
lineages with recent losses (Ramesh et al. 2005). 2008) and once per 50,000 asexual cycles for Saccharomyces

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


The q/hw ratio was lower in P. tapetis (table 3) than in cerevisiae (Ruderfer et al. 2006). Some protists have an ob-
S. paradoxus (0.31–0.81) (Tsai et al. 2008) or Plasmodium ligate sexual stage because of constraints such as morphol-
falciparum (0–6.2) (Awadella 2003) but greater than that ogy (e.g., diatom frustule size), development (alternation of
in Trypanosoma cruzi, a predominantly clonal species generations in foraminiferans), or host relationships (e.g.,
(Llewellyn et al. 2009) with a q/hw of zero (Awadella some apicomplexans). Even given our wide range of param-
2003). In the animals Drosophila melanogaster and Caeno- eters, if P. tapetis had a similar obligate sexual cycle, we
rhabditis remanei, the ratios ranged from 0.6 to 7.6 (Tsai would have expected to see a lower bound for the number
et al. 2008). Fraser et al. (2007) estimated that q/hw ratios of asexual cycles per sexual cycle.
greater than 0.25 prevented clonal divergence in bacteria,
which rely on homologous recombination for sex. Uniformly Host Relationship and Demographics
low values, below 0.07 in three different methods of Having isolated P. tapetis 126 times from bivalve guts, we
analysis, suggested that clonality has been an important force suspect that guts are its main habitat. Anurofeca richardsi,
in shaping the genetic structure in our P. tapetis population. the closest relative to P. tapetis with any in situ observa-
This superimposition of clonality on sexuality is often ob- tions, grows and divides in tadpole digestive tracts (Beebee
served in populations of unicells (e.g., Ngouanesavanh and Wong 1993). P. tapetis is unlikely to be as abundant in
et al. 2006; Bui et al. 2008). the water column as it is within the digestive tract because,
in repeated attempts, we never isolated it from the undi-
Frequency of Sex luted mantle cavity fluid of bivalves even when it could be
Our estimates of the ratio of sexual to asexual cycles in P. cultured from the guts of the same animal. It is not known
tapetis were based on a series of assumptions. One assump- how long P. tapetis resides in its host’s digestive tract, nor
tion was that no inbreeding was occurring. However, if in- has its growth there been physically observed. That P. ta-
breeding does occur, as is likely the case, our calculations petis may have just been ingested by bivalves as food is pos-
will underestimate mating frequency. The higher the fre- sible but food can pass through a bivalve digestive tract
quency of inbreeding, the larger our underestimate of mat- within a day (Ren et al. 2006), and the high frequency
ing frequency would be. Inbreeding coefficients are of isolation of P. tapetis would then have to be explained
unknown for P. tapetis but using Ne 5 q/2r (1  F) and by a high incidence in the water column. To reproduce suc-
Ne 5 h (1 þ F)/2l where F is the inbreeding coefficient (Tsai cessfully in the gut P. tapetis would need a mechanism for
et al. 2008), assuming that F 5 0.5 would increase the q/hw retention beyond the day needed for passage of food be-
ratios and mating frequency estimates by ;1.33 times, cause, based on its growth in culture, P. tapetis endospores
whereas F 5 0.9 would increase them by ;5 times. took 3–5 days to grow to the reproductive stage. Con-
Our calculations of the frequency of mating are based on versely, P. tapetis’s variation in abundance among even co-
the relative frequency of meiotic compared with mitotic habiting hosts suggests that it comes and goes rather than
divisions, equating a nuclear mitotic cycle to an asexual accumulating continuously over the lifespan of the host
generation. We could alternatively define a single asexual (table 1).
generation as the time between initiation of endospore According to our estimates of the frequency of sex, in-
growth and spore release (fig. 4). In culture, endospore fection of a host animal does not depend on a sexual stage.
growth involved five to nine rounds of mitosis, resulting If P. tapetis undergoes about two mitotic divisions per day
in sporangia with between 32 and 512 uninucleate endo- in the host as it does in culture, it can undergo approxi-
spores. In nature, the number of mitotic divisions before mately 700 mitotic divisions per year. Assuming an average
endospore release remains unknown, but assuming six retention time of a year and our median estimates of

2023
Marshall and Berbee · doi:10.1093/molbev/msq078 MBE
frequency of sex, P. tapetis must pass through several hosts mals, and if other unicellular opisthokonts are haploid as
before mating. If sex was obligate and infection limited to well, then haploid dominant life cycles may have been the
sexually derived resting spores or zygotes, the retention ancestral condition of the metazoan stem lineage.
time necessary before sexual reproduction could equal
or surpass the host lifespan (fig. 6). Supplementary Material
The average level of silent-site pairwise nucleotide diver-
sity (hw) was within the range but lower than the average Supplementary tables S1 and S2 are available at Molecular Bi-
value of 0.057 of other unicellular eukaryotes reported by ology and Evolution online (http://www.mbe.oxfordjournals
Lynch (2006). The effective population size of P. tapetis was .org/).
on the order of 106 to 107, which is typical for unicellular
eukaryotes (Lynch 2006). Effective population sizes are Acknowledgments
based on numbers of individuals expected to contribute We thank S. Otto for helpful discussions on coalescent the-
progeny to the next generation. Because many cells die ory, W. Maddison for Poisson probability calculations, and
and do not contribute to the gene pool, the effective size L. Rieseberg and several anonymous reviewers for com-
is lower than the censused population. Pseudoperkinsus ta- ments on earlier versions of this manuscript. This work
petis can grow and divide for a few days in unsupplemented was supported by the Natural Sciences and Research Coun-
seawater but cells arrest at uni- and binucleate stages. cil of Canada and the Bamfield Marine Sciences Centre.
Reinfection is probably a limiting factor in P. tapetis’s de-

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


mographics. The envisioned challenges of colonization and References
reinfection faced by symbionts such as P. tapetis may be Adl S, Simpson AG, Farmer MA, et al. (28 co-authors). 2005. The
offset by factors, such as reduced competition, ample sup- new higher level classification of eukaryotes with emphasis on
ply of nutrients, or the ability to proliferate and mate once the taxonomy of protists. J Eukaryot Microbiol. 52:399–451.
past the reinfection hurdle. Agapow PM, Burt A. 2001. Indices of multilocus linkage
Our finding of a large population with evidence of re- disequilibrium. Mol Ecol Notes. 1:101–102.
Archie JW. 1989. A randomization test for phylogenetic information
combination suggests that P. tapetis is a mostly overlooked in systematic data. Syst Zool. 38:239–252.
endemic of a wide host range of shellfish and other inver- Awadella P. 2003. The evolutionary genomics of pathogen re-
tebrates rather than a recent colonizer. Nucleotide diversity combination. Nat Rev Genet. 4:50–60.
and effective population size decrease during population Beebee TJC, Wong ALC. 1993. Stimulation of cell division and DNA
bottlenecks such as recent colonization (Smith 1991; Rich replication in Prototheca richardsi by passage through larval
and Ayala 2000). The presence of a recombining population amphibian guts. Parasitology 107:119–123.
also suggests a natural ecological niche (Campbell et al. Bernander R, Palm JED, Svärd SG. 2001. Genome ploidy in different
stages of the Giardia lamblia life cycle. Cell Microbiol. 3:55–62.
2005). For example, recombination in Giardia was shown Brondani C, Brondani RPV, Garrido LR, Ferreira ME. 2000.
once isolates were collected from an endemic area with Development of microsatellite markers for the genetic analysis
high transmission rates and mixed infections (Cooper of Magnaporthe grisea. Genet Mol Biol. 23:753–762.
et al. 2007). Although most frequently isolated from bi- Brown AHD, Feldman MW, Nevo E. 1980. Multilocus structure of
valves (Figueras et al. 2000; Takishita et al. 2008), Pseudo- natural populations of Hordeum spontaneum. Genetics 96:523–536.
perkinsus showed no obvious detriment to its hosts. We Bruen T, Philippe H, Bryant D. 2006. A simple and robust statistical
found no evidence of host specificity based on the distri- test for detecting the presence of recombination. Genetics
172:2665–2681.
bution of haplotypes.
Bui T, Lin X, Malik R, Heitman J, Carter D. 2008. Isolates of
Cryptococcus neoformans from infected animals reveal genetic
exchange in unisexual, a mating type populations. Eukaryot Cell.
Conclusions 7:1771–1780.
We have demonstrated for the first time among any mem- Burt A, Carter DA, Koenig GL, White TJ, Taylor JW. 1996. Molecular
ber of the basal unicellular opisthokont lineages that sexual markers reveal cryptic sex in human pathogen Coccidiodes
recombination occurs and that sex is most likely faculta- immitis. Proc Natl Acad Sci U S A. 93:770–773.
tive. According to our estimates of a low frequency of Campbell LT, Carter D. 2006. Looking for sex in the fungal
sex, which were based on a small amount of data, P. tapetis pathogens Cryptococcus neoformansand Cryptococcus gattii.
FEMS Yeast Res. 6:588–598.
most likely enters the host as an asexually derived endo- Campbell LT, Currie BJ, Krockenberger M, Malik R, Meyer W,
spore. Ichthyosporeans are one of few host-dependent Heitman J, Carter D. 2005. Clonality and recombination in
lineages among the opisthokonts. Phylogenetically, multi- genetically differentiated subgroups of Cryptococcus gattii.
cellular animals originated after the ichthyosporean lineage Eukaryot Cell. 4:1403–1409.
diverged and so the earliest ichthyosporeans could not Carr M, Nelson M, Leadbeater BSC, Baldauf SL. 2008. Three families
have had metazoan hosts. The lack of host specificity in of LTR retrotransposons are present in the genome of the
P. tapetis may reflect the relatively recent evolution of ich- choanoflagellate Monosiga brevicollis. Protist 159:579–590.
Cavalier-Smith T. 1987. The origin of fungi and pseudofungi. In:
thyosporean dependence on metazoan hosts. Our data are
Rayner A, Brasier C, Moore D, editors. Evolutionary biology of
the first to reveal ploidy in any member of the unicellular fungi. Cambridge: Cambridge University Press. p. 339–353.
relatives of animals. A potentially haploid lineage at the Clark A. 1990. Inference of haplotypes from PCR-amplified samples
base of the metazoan tree contrasts with the diploid ani- of diploid populations. Mol Biol Evol. 7:111–122.

2024
Cryptic Sex and Life History of Pseudoperkinsus · doi:10.1093/molbev/msq078 MBE
Clement M, Posada D, Crandall KA. 2000. TCS: a computer program nant sequences and recombination breakpoints. AIDS Res Hum
to estimate gene genealogies. Mol Evol. 9:1657–1659. Retroviruses. 21:98–102.
Cooper MA, Adam RD, Worobey M, Sterling CR. 2007. Population Martin D, Williamson C, Posada D. 2005. RDP2: recombination
genetics provides evidence for recombination in Giardia. Curr detection and analysis from sequence alignments. Bioinformatics
Biol. 17:1984–1988. 21:260–262.
Cunningham CW. 1997. Can three incongruence tests predict when Maynard Smith J, Smith NH, O’Rourke M, Spratt BG. 1993. How
data should be combined? Mol Biol Evol. 14:733–740. clonal are bacteria? Proc Natl Acad Sci U S A. 90:4384–4388.
Farris JS, Kallersjo M, Kluge AG, Bult C. 1994. Testing significance of McVean G, Awadella P, Fearnhead P. 2002. A coalescent-based
incongruence. Cladistics 10:315–319. method for detecting and estimating recombination from gene
Figueras A, Lorenzo G, Ordas MC, Gouy M, Novoa B. 2000. Sequence sequences. Genetics 160:1231–1241.
of the small subunit ribosomal DNA gene of Pseudoperkinsus Mendoza L, Taylor JW, Ajello L. 2002. The class Mesomycetozoea:
tapetis isolated carpet shell clam in Galicia (NW of Spain). Mar a heterogeneous group of microorganisms at the animal-fungal
Biotechnol. 2:419–428. boundary. Annu Rev Microbiol. 56:315–344.
Fraser C, Hanage WP, Spratt BG. 2007. Recombination and the Ngouanesavanh T, Guyot K, Certad G, Le Fichoux Y, Chartier C,
nature of bacterial speciation. Science 315:476–480. Verdier RI, Cailliez JC, Camus D, Dei-Cas E, Bañuls AL. 2006.
Geiser DM, Pitt JI, Taylor JW. 1998. Cryptic speciation and Cryptosporidium population genetics: evidence of clonality
recombination in the aflatoxin-producing fungus Aspergillus in isolates from France and Haiti. J Eukaryot Microbiol. 53:
flavus. Proc Natl Acad Sci U S A. 95:388–393. S33–S36.
Hibbits J. 1978. Marine Eccrinales (Trichomycetes) found in Oliveira RP, Broude NE, Macedo AM, Cantor CR, Smith CL,
crustaceans of the San Juan Archipelago, Washington. Syesis Pena SDJ. 1998. Probing the population genetic structure of

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012


11:213–261. Trypanosoma cruzi with polymorphic microsatellites. Proc Natl
Holmes EC, Worobey M, Rambaut A. 1999. Phylogenetic evidence Acad Sci U S A. 95:3776–3780.
for recombination in dengue virus. Mol Biol Evol. 16:405–409. Ramesh MA, Malik SB, Logsdon JMJ. 2005. A phylogenomic
Hudson RR. 1987. Estimating the recombination parameter of a finite inventory of meiotic genes: evidence for sex in Giardia and an
population model without selection. Genet Res. 50:245–250. early eukaryotic origin of meiosis. Curr Biol. 15:185–191.
Hudson RR, Kaplan N. 1985. Statistical properties of the number of Ren JS, Ross AH, Hayden BJ. 2006. Comparison of assimilation
recombination events in the history of a sample of DNA efficiency on diets of nine phytoplankton species of the
sequences. Genetics 111:147–164.
greenshell mussel Perna canaliculus. J Shellfish Res. 25:887–892.
Huson DH. 1998. SplitsTree: a program for analyzing and visualizing
Reynolds DR. 1993. The fungal holomorph: an overview. In: Reynolds
evolutionary data. Bioinformatics 14:68–73.
DR, Taylor JW, editors. The fungal holomorph: a consideration of
Huson DH, Bryant D. 2006. Application of phylogenetic networks in
mitotic and pleomorphic speciation. Wallingford (United
evolutionary studies. Mol Biol Evol. 23:254–267.
Kingdom): CAB International. p. 15–26.
Huson DH, Kloepper TH. 2005. Computing recombination networks
Rich SM, Ayala FJ. 2000. Population structure and recent evolution
from binary sequences. Bioinformatics 21:ii159–ii165.
of Plasmodium falciparum. Proc Natl Acad Sci U S A.
Johannsen E, van der Walt JP. 1980. Hybridization studies within
97:6994–7001.
the genus Saccharomyces Klocker. Can J Microbiol. 26:
Rozas J, Sanchez-DelBarrio JC, Messeguer X, Rozas R. 2003. DnaSP,
1199–1203.
Johnson LJ, Koufopanou V, Goddard MR, Hetherington R, DNA polymorphism analyses by the coalescent and other
Schäfer SM, Burt A. 2003. Population genetics of the wild yeast methods. Bioinformatics 19:2496–2497.
Saccharomyces paradoxus. Genetics 66:43–52. Ruderfer DM, Pratt SC, Seidel HS, Kruglyak L. 2006. Population
Jostensen JP, Sperstad S, Johansen S, Landfald B. 2002. Molecular- genomic analysis of outcrossing and recombination in yeast. Nat
phylogenetic, structural and biochemical features of a cold Genet. 38:1077–1081.
adapted, marine ichthyosporean near the animal-fungal di- Ruiz-Trillo I, Lane CE, Archibald JM, Roger AJ. 2006. Insights into the
vergence, described from in vitro cultures. Eur J Protistol. evolutionary origin and genome architecture of the unicellular
38:93–104. opisthokonts Capsaspora owczarzaki and Sphaeroforma arctica.
Llewellyn MS, Miles MA, Carrasco HJ, et al. (11 co-authors). 2009. J Eukaryot Microbiol. 53:1–6.
Genome-scale multilocus microsatellite typing of Trypanosoma Ruiz-Trillo I, Roger AJ, Burger G, Gray MW, Lang BF. 2008. A
cruzidiscrete typing unit 1 reveals phylogeographic structure and phylogenomic investigation into the origin on metazoa. Mol Biol
specific genotypes linked to human infection. PLoS Pathog. Evol. 25:664–672.
5:e1000410. Rynearson TA, Armbrust EV. 2000. DNA fingerprinting reveals
LoBuglio KF, Taylor JW. 2002. Recombination and genetic extensive genetic diversity in a field population of the centric
differentiation in the mycorrhizal fungus Cenococcum geophilum diatom Ditylum brightwelli. Limnol Oceanogr. 45:1329–1340.
Fr. Mycologia. 94:772–780. Santos SR, Coffroth MA. 2003. Molecular genetic evidence that
Lynch M. 2006. The origins of eukaryotic gene structure. Mol Biol dinoflagellates belonging to the genus Symbiodinium Freuden-
Evol. 23:450–468. thal are haploid. Biol Bull. 204:10–20.
Maddison WP, Maddison DR. 2000. MacClade 4: interactive analysis Seifert KA, Samson RA, deWaard JR, Houbraken J, Levesque CA,
of phylogeny and character evolution. Sunderland (MA): Moncalvo J-M, Louis-Seize G, Hebert PDN. 2007. Prospects for
Sinauer. fungus identification using C01 DNA barcodes, with Penicillium
Maldonado M. 2005. Choanoflagellates, choanocytes, and animal as test case. Proc Natl Acad Sci U S A. 104:3901–3906.
multicellularity. Invert Biol. 123:1–22. Shalchian-Tabrizi K, Minge MA, Espelund M, Orr R, Ruden T,
Marshall WL, Celio G, McLaughlin DJ, Berbee ML. 2008. Multiple Jakobsen KS, Cavalier-Smith T. 2008. Multigene phylogeny of
isolations of a culturable, motile Ichthyosporean (Mesomyceto- choanozoa and the origin of animals. PLoS ONE. 3:e2098.
zoa, Opisthokonta), Creolimax fragrantissima n. gen., n. sp., from Smith JM. 1991. The population genetics of bacteria. Proc Biol Sci.
marine invertebrate digestive tracts. Protist. 159:415–433. 245:37–41.
Martin D, Posada D, Crandall KA, Williamson C. 2005. A modified Steenkamp ET, Wright J, Baldauf SL. 2006. The protistan origins of
bootscan algorithm for automated identification of recombi- animals and fungi. Mol Biol Evol. 23:93–106.

2025
Marshall and Berbee · doi:10.1093/molbev/msq078 MBE
Swofford DL. 1998. PAUP**: phylogenetic analysis using parsimony Tsai IJ, Bensasson D, Burt A, Koufopanou V. 2008. Population
(*and other methods). Version 4. Sunderland (MA): Sinauer genomics of the wild yeast Saccharomyces paradoxus: quanti-
Associates. fying the life cycle. Proc Natl Acad Sci U S A. 105:4957–4962.
Takishita K, Fujiwara Y, Kawato M, Kakizoe N, Miyazaki M, van der Verde M, During HJ, van de Zande L, Bijlsma R. 2001. The
Maruyama T. 2008. Molecular identification of the ichthyospor- reproductive biology of Polytrichum formosum clonal structure
ean protist ‘‘Pseudoperkinsus tapetis’’ from mytilid mussel and paternity revealed by microsatellites. Mol Ecol. 10:2423–2434.
Adipicola pacifica associated with submerged whale carcasses Wakeley J. 1997. Using the variance of pairwise differences to
in Japan. Mar Biotechnol. 10:13–18. estimate the recombination rate. Genet Res. 69:45–48.
Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG. Watterson GA. 1975. On the number of segregating sites in
1997. The Clustal X windows interface flexible strategies for genetical models without recombination. Theor Popul Biol. 7:
multiple sequence alignment aided by quality analysis tools. 256–276.
Nucleic Acids Res. 24:4876–4882. Weeks AR, Marec F, Breeuwer JAJ. 2001. A mite species that consists
Tibayrenc M, Kjellberg F, Ayala FJ. 1990. A clonal theory of parasitic entirely of haploid females. Science 292:2479–2482.
protozoa: the population structures of Entamoeba, Giardia, Whisler HC. 1968. Developmental control of Amoebidium para-
Leishmania, Naegleria, Plasmodium, Trichomonas, and Trypano- siticum. Dev Biol. 17:562–570.
soma and their medical and taxonomical consequences. Proc Xu J. 2004. The advantage of sex in evolving yeast populations.
Natl Acad Sci U S A. 87:2414–2418. Nature 388:465–468.

Downloaded from http://mbe.oxfordjournals.org/ by guest on September 15, 2012

2026

You might also like