ArticlePDF Available

Tubisorus, a new genus of smut fungi (Ustilaginomycetes) for Sorosporium pachycarpum

Authors:
129
MYCOLOGIA BALCANICA 8: 129–135 (2011)
Tubisorus, a new genus of smut fungi (Ustilaginomycetes) for
Sorosporium pachycarpum
Kálmán Vánky ¹* & Matthias Lutz ²
¹ Herbarium Ustilaginales Vánky (H.U.V.), Gabriel-Biel-Str. 5, D-72076 Tübingen, Germany
² Organismische Botanik, Institut für Evolution und Ökologie, Universität Tübingen, Auf der Morgenstelle 1, D-72076
Tübingen, Germany
Received 22 October 2011 / Accepted 21 November 2011
Abstract. A new genus of smut fungi, Tubisorus, is proposed for Sorosporium pachycarpum on Mnesithea
rottboellioides, Poaceae.
Key words: molecular analysis, new combination, new genus, smut fungi, taxonomy, Tubisorus, T. pachycarpus
*Corresponding author: e-mail: vanky.k@cityinfonetz.de
Introduction
A peculiar smut fungus, Sorosporium pachycarpum was
described by H. Sydow (in Sydow & Petrak 1928) on
Rottboellia ophiuroides (= Mnesithea rottboellioides), from the
Philippines.  e sori are very long, tubiform, in the axis of
spikes with aborted spikelets.  ey are  lled with a black,
granular powdery mass of spore balls.  e spore balls are
permanent, composed of globoid, thick-walled, ornamented
spores.  e fungus was placed into the genus Tolyposporella
by Ling (1949), and into the genus Endosporisorium by
Vánky (1995); the latter genus turned out to be a synonym of
Macalpinomyces. irumalachar et al. (1967: 395) treated it
also under Tolyposporella. Vánky (1998: 93–95) considered the
genus Sorosporium F. Rudolphi to be a synonym of ecaphora
Fingerh., parasitising only dicotyledonous host plants.
Based on the peculiar sorus, spore ball and spore
morphology, as well as on molecular phylogenetic analyses of
ITS and LSU rDNA sequences, a new genus is proposed for
S. pachycarpum.
Materials and methods
e following specimens were examined for the present work:
Philippines, Luzon, Pampanga Prov., Stotsenberg, XI.1923,
coll. M.S. Clemens 2313 (BPI 180089; syntype); Papua
New Guinea, Morobe, 19.XI.1939, coll. M.S. Clemens
(BPI 71558, H.U.V. 10504 & 16772); PNG, Moresby,
20.III.1998, coll. R.G. Shivas, G.R. Kula, C. & K. Vánky, in
Vánky, Ust. exs. No. 1036 (H.U.V. 18515 & 18606); PNG,
Port Moresby, 21.III.1998, coll. R.G. Shivas, C. & K. Vánky
(H.U.V. 18520); PNG, Finschafen, 31.III.1998, coll. R.G.
Shivas, G.R. Kula, C. & K. Vánky (H.U.V. 18521); Australia,
Queensland, Townsville, James Cooke Univ. Campus, coll.
C. Bretzel, IX.1996 (H.U.V. 18206); Queensland, Cape York
Peninsula, 3 km N of Bamaga, coll. R.G. Shivas, C. & K.
Vánky (H.U.V. 19237); Queensland, 65 km S of Cairns,
Eubenangee Swamp, 17.VIII.2006, coll. R.G. Shivas, C. &
K. Vánky (H.U.V. 21582); Australia, Northern Territory,
between Pine Creek and Adelaide River, 12.4 km o Stuart
Hwy, Dorat Road, coll. M. & R.G. Shivas, Y.C. Tran, T. & K.
Vánky, 23.IV.2011 (H.U.V. 21891).
e nomenclatural novelty was registered in MycoBank
(www.MycoBank.org, Crous et al. 2004).
Morphological examination
Sorus structure and spore characteristics were studied using
dried herbarium specimens. For microscopic studies of
the soral, spore ball and spore characters, young sori were
rehydrated in hot water,  xed in 2 % glutaraldehyde in 0.1
M Na-cacodylate bu er at pH 7.2 for several days. After six
transfers in 0.1 M Na-cacodylate bu er, the material was
post xed in 1 % osmiumtetroxide in the same bu er for 1 h
130 , .  , . — Tubisorus,       (Ustilaginomycetes)  Sorosporium pachycarpum
in the dark, washed in distilled water, and stained in 1 %
aqueous uranyl acetate for 1 h in the dark. After  ve washes in
distilled water, the material was dehydrated in acetone series,
embedded in Spurr’s plastic and sectioned with a diamond
knife. Semi-thin sections were stained with new fuchsin and
crystal violet, mounted in “Entellan” and studied in a light
microscope. For light microscopy (LM) spore balls were
dispersed in a small droplet of lactophenol, covered with a
cover glass, gently heated to boiling point to rehydrate the
spores and expel air bubbles from the preparation, and studied
at 1000× magni cation. For scanning electron microscopy
(SEM), spore balls were placed on double-sided adhesive
tape, mounted on a specimen stub, sputter-coated with gold-
palladium, ca 35 nm, and examined in a SEM at 10 kV.
DNA extraction, PCR, and sequencing
Genomic DNA was isolated directly from the herbarium
specimen (H.U.V. 21891). For methods of isolation and
crushing of fungal material, DNA extraction, ampli cation,
puri cation of PCR products, sequencing, and processing of
the raw data see Lutz et al. (2004). ITS 1 and ITS 2 regions
of the rDNA including the 5.8S rDNA (ITS) were ampli ed
using the primer pair ITS1-F (Gardes & Bruns 1993) and
ITS4 (White et al. 1990).  e 5´-end of the nuclear large
subunit ribosomal DNA (LSU) was ampli ed using the
primer pair NL1 and NL4 (O’Donnell 1993). Primers were
used for both PCR and cycle sequencing. For ampli cation
the annealing temperature was adjusted to 48 °C. DNA
sequences determined in this study were deposited in GenBank
(accession numbers: ITS: JN871718/LSU: JN871717).
Phylogenetic analyses
Blast searches (Altschul et al. 1997) for both the ITS and
LSU sequence revealed closest similarity to members of the
Macalpinomyces/Sporisorium/Ustilago-group. To further elu-
cidate the phylogenetic position of Sorosporium pachy carpum,
its ITS and LSU sequence was analysed against the combined
ITS/LSU-dataset of Stoll et al. (2005). For this analysis, the
dataset was reduced to one specimen per species; the ITS and
LSU sequences of Anomalomyces panici (Vánky et al. 2006)
were added. GenBank accession numbers of sequences used
in the molecular analyses are included in Fig. 8.
Sequence alignment was obtained using MAFFT 6.853
(Katoh et al. 2002, 2005; Katoh & Toh 2008) using the
L-INS-i option. As suggested by Giribet & Wheeler (1999)
and Gatesy et al. (1993), respectively, to obtain reproducible
results manipulation of the alignment by hand as well as
manual exclusion of ambiguous sites were avoided. Instead,
highly divergent portions of the alignment were omitted
using GBlocks 0.91b (Castresana 2000) with the following
options: ‘Minimum Number of Sequences for a Conserved
Position’ to 50, ‘Minimum Number of Sequences for a Flank
Fig. 1. Tubisorus pachycarpus on Mnesithea rottboellioides (Pa-
pua New Guinea, H.U.V. 16 772) – young sori in much elon-
gate, tubular axis of the spikes, with aborted spikelets. Hab-
it. Bar = 1 cm
Position’ to 50, ‘Maximum Number of Contiguous Non-
conserved Positions’ to 8, ‘Minimum Length of a Block’ to 5
and ‘Allowed Gap Positions’ to ‘With half’.
e resulting alignment [new number of positions:
1160 (65 % of the original 1779 positions) number of
variable sites: 372] was used for phylogenetic analyses using
a Bayesian Approach (BA) and Maximum Likelihood (ML).
For BA a Markov chain Monte Carlo technique was used
as implemented in the computer program MrBayes 3.1.2
(Huelsenbeck & Ronquist 2001; Ronquist & Huelsenbeck
   () 131
2003). Four incrementally heated simultaneous Markov
chains were run over 5 000 000 generations using the general
time reversible model of DNA substitution with gamma
distributed substitution rates and estimation of invariant sites,
random starting trees and default starting parameters of the
DNA substitution model as recommended by Huelsenbeck &
Rannala (2004). Trees were sampled every 100th generation,
resulting in an overall sampling of 50 001 trees. From these,
the  rst 5 001 trees were discarded (burnin = 5 001).  e trees
sampled after the process had reached stationarity (45 000
trees) were used to compute a 75 % majority rule consensus
tree to obtain estimates for the a posteriori probabilities of
groups of species.  is Bayesian approach to phylogenetic
analysis was repeated  ve times to test the independence of
the results from topological priors (Huelsenbeck et al. 2002).
ML analysis (Felsenstein 1981) was conducted with the
RAxML 7.2.8 software (Stamatakis 2006), using raxmlGUI
(Silvestro & Michalak 2010), invoking the GTRCAT and the
rapid bootstrap option (Stamatakis et al. 2008) with 1000
replicates.
In line with Stoll et al. (2005), trees were rooted with
sequences of Eriomoeszia eriocauli and Moesziomyces bullatus.
Results
Phylogenetic analyses
e di erent runs of BA that were performed and the ML
analyses yielded consistent topologies in respect to well
supported branchings. To illustrate the results, the consensus
tree of one run of the Bayesian phylogenetic analyses is
presented (Fig. 8). Estimates for a posteriori probabilities are
indicated on branches before slashes, numbers on branches
after slashes are ML bootstrap support values.
In all analyses Sorosporium pachycarpum clustered within
the Macalpinomyces/Sporisorium/Ustilago-group as sister taxon
of Ustilago bouriquetii and U. maydis within a group consisting
of Macalpinomyces loudetiae, M. simplex, M. trichopterygis,
M. tristachyae, Sporisorium trachypogonis-plumosi, and U.
vetiveriae.
Taxonomy
Tubisorus Vánky & M. Lutz, gen. nov.
MB  563575
Sori formationes tubulares plantarum nutrientium familiae
Poacearum. Massa atra granuloso-pulverea glomerulorum
sporarum conferti, columella et cellulae steriles nullae. Glomeruli
sporarum e sporis tantum compositi. Sporae pigmentatae
(brunneae, sine violaceo nec rubro tincto). Sporae verae
hyalinae, in matrice carassa, gelatinosa, hyalina immittae, strato
externo tenui, pigmentato, ornamentis velatis.
Typus generis: T. pachycarpus.
Sori tubular on host plants in the Poaceae, lled with dark,
granular powdery mass of spore balls, columella and sterile
cells lacking. Spore balls composed of spores only. Spores
pigmented (brown, without violet or red tint).  e proper
spore is hyaline and embedded in a thick, gelatinous, hyaline
substance, coated by a thin, pigmented, ornamented outer
layer of spore wall.
Type of the genus: T. pachycarpus.
Tubisorus pachycarpus (Syd.) Vánky & M. Lutz, comb. nov.
MB  563576
Basionym: Sorosporium pachycarpum Syd., in Sydow &
Petrak, Ann. Mycol. 26: 431, 1928; Tolyposporella pachycarpa
(Syd.) L. Ling, Sydowia 3: 133, 1949; Endosporisorium
pachycarpum (Syd.) Vánky, Mycotaxon 56: 213, 1995. –
Syntype on Rottboellia ophiuroides (= Mnesithea rottboellioides),
Philippines, Luzon, Pampanga Prov., Stotsenberg, XI.1923,
coll. M.S. Clemens 2313 (BPI 180089!) (another syntype is
probably lost).
Sori (Figs 1, 7) in much elongated axis of the spikes with
aborted spikelets, tubular, 0.5–1.5 mm wide, up to 20–25
cm long, yellowish to greyish brown, at maturity rupture
longitudinally, on several places between the veins, exposing
the black, granular-powdery mass of spore balls. Spore balls
(Figs 2–5) globoid, long ellipsoidal to irregular, composed of
5–50 or more spores, apparently loose but rather permanently
connected, 25–50 (–65) × 45–100 (–150) µm, dark olivaceous
brown. Spores (Figs 3–5) subglobose, ovoid or irregular with
one or several  attened contact sides, 12–20 × 16–24 µm,
olivaceous brown; wall 3-layered (Figs 3, 5): outer layer 1.5–
2 µm thick, pigmented, densely provided with irregular,
prismatic warts, c. 1 µm high on the free surface, smaller and
ner on the contact sides; middle layer hyaline, gelatinous,
homogenous, 2.5–7 µm thick; inner layer 0.5–1.5 µm thick,
smooth, surrounding the  nely granular, hyaline proper
spores which in squashed spores are eliberated (Fig. 6).  ese
are subglobose, broadly ellipsoidal, ovoid, subpolyhedrally
irregular or elongated, (7–) 8–10.5 × 9–14.5 µm; wall 0.5–
1 µm thick, composed of a very thin inner layer and and a
hyaline sheath, smooth. Spore germination unknown.
On Poaceae: Mnesithea rottboellioides (R. Br.) de Koning &
Sosef (Coelorachis rottboellioides (R. Br.) A. Camus; Manisuris
rottboellioides (R. Br.) Kuntze; Rottboellia ophiuroides Benth.).
Geographic distribution: Australia, Papua New Guinea,
Philippines (comp. Sydow & Petrak 1928: 431; Shivas et al.
2001: 335; Vánky & Shivas 2008: 105).
Discussion
H. Sydow (in Sydow & Petrak 1928: 431) described
Sorosporium pachycarpum to be in the leaves of the host plant.
Ling (1949: 133) also considered that the sori are in the leaves,
which was the reason that he transferred the fungus into the
genus Tolyposporella. Zundel (1953: 68) described the sori on
the abaxial side of the leaves, and treated the fungus under
132 , .  , . — Tubisorus,       (Ustilaginomycetes)  Sorosporium pachycarpum
Figs 2–4. Tubisorus pachycarpus on Mnesithea rottboellioides (Papua New Guinea, H.U.V. 16772). 2. Semi-thin, stained, transversal
section of a part of a sorus with sectioned spore balls. Bar = 25 µm. 3–4. Spores balls and spores in LM and in SEM. Bars = 10 µm
the genus Sorosporium. Despite the fact that  irumalachar
et al. (1967: 395) wrote that “Many Sorosporium species are
foliicolous . . . . ” and that “ e smut was correctly placed
in the genus Sorosporium by Sydow”, they treated it as
Tolyposporella pachycarpa (Syd.) Ling. Vánky & Shivas (2008:
166) wrote that the sori of Tolyposporella pachycarpa are “in
the axis of aborted in orescence”, which is inaccurate.  e
exact place of the sori is  rst described by Vánky (2012: 995).
e type of the genus Tolyposporella G.F. Atk. is T.
chrysopogonuis G.F. Atkinson (1897) on Sorghastrum nutans
(L.) Nash, USA. Its sori form linear, more or less con uent,
initially subepidermal, later bursting striae on the inner
surface of leaf sheaths.  e spore balls are composed of  rmly
agglutinated, smooth spores with unevenly thickened wall. It is
quite di erent from the smut on Mnesithea. Tolyposporella is a
genus of  ve species, four on Poaceae, and one on Eriocaulaceae,
which probably does not belong to this genus (comp. Vánky
2012). All Tolyposporella species have spores with smooth,
unevenly thick, multilamellar wall and sori on the surface
of the leaves or leaf sheaths. In contrast, the spore wall of
Tubisporus pachycarpus is ornamented and three-layered (for
description see above). Similar spore wall structure, in which
   () 133
Figs 5–6. Tubisorus pachycarpus on Mnesithea rottboellioides (Pa-
pua New Guinea, H.U.V. 16772). 5. Sectioned and stained
spore balls and spores showing the thin, pigmented, ornament-
ed outer spore wall, the thick, homogenous mass in which the
thin-walled proper spores with granular content are embed-
ded. Bar = 10 µm. 6. Ruptured outer, pigmented spore walls
permitting the thin-walled, hyaline proper spores to emerge.
Bar = 10 µm
the proper spore is embedded in a gelatinous layer and coated
by a thin, pigmented outer layer occurs in the smut fungi only
in the genus Kuntzeomyces Henn. ex Sacc., with two closely
related species in the spikelets of Rhynchospora (Cyperaceae;
comp. Piepenbring 2001). In them, however, the spores are
smooth and single.  ese characters of the spores, in addition
to the soral characters, exclude Tubisorus pachycarpus from
Tolyposporella. Tubisorus neither belongs to the genus Spori-
sorium Ehrenb. ex Link, nor to Macalpinomyces Langdon &
Full., in which sterile cells, or groups of sterile cells, and in the
sori often also columella/ae are present (comp. Vánky 2002).
Piepenbring et al. (1998: 210, Fig. 28) reported presence
of germ pores in the spores of “Sorosporiumpachycarpum,
which could not be con rmed.
Molecular phylogenetic analyses based on BA and
ML analyses of ITS + LSU rDNA sequence data suggest a
Fig. 7. Mature sori of Tubisorus pachycarpus on Mnesithea rot-
tboellioides (Photo: Christine Vánky)
common ancestor of Tubisorus pachycarpus and the Ustilago
maydis/U. bouriquetii clade. U. maydis (DC.) Corda produces
galls on the stems, leaves or in orescence of Zea (including
Euchlaena) species (subtribe Tripsacinae, tribe Andropogoneae,
subfam. Panicoideae),  lled with single spores. U. bouriquetii
Maubl. & Roger produces considerably hypertrophied
laments of Stenotaphrum species (subtribe Setariinae, tribe
Paniceae, subfam. Panicoideae),  lled by single spores.
Mnesithea Kunth belongs to the subtribe Rottboelliinae
of the tribe Andropogoneae, subfam. Panicoideae (Clayton
& Renvoize 1986: 368).  ree other known smut fungi on
members of this genus are: Sporisorium operculatum Vánky
& R.G. Shivas (2001: 154; Australia), S. perforatum Vánky
(2004: 104; India), and S. mnesitheae (Mishra) Vánky (2004:
107; India). None of these, including Ustilago maydis and U.
bouriquetii is similar to Tubisorus pachycarpus.
134 , .  , . — Tubisorus,       (Ustilaginomycetes)  Sorosporium pachycarpum
Fig. 8. Bayesian inference of
phylogenetic position of So-
rosporium pachycarpum: Markov
chain Monte Carlo analysis of an
alignment of concatenated ITS
+ LSU base sequences using the
GTR+I+G model of DNA sub-
stitution with gamma distribut-
ed substitution rates and estima-
tion of invariant sites, random
starting trees and default start-
ing parameters of the DNA sub-
stitution model. A 75 % major-
ity-rule consensus tree is shown
computed from 45 000 trees that
were sampled after the process
had reached stationarity.  e to-
pology was rooted with sequenc-
es of Eriomoeszia eriocauli and
Moesziomyces bullatus. Numbers
on branches before slashes are
estimates for a posteriori prob-
abilities, numbers on branches
after slashes are ML bootstrap
support values. Branch lengths
were averaged over the sampled
trees.  ey are scaled in terms
of expected numbers of nucle-
otide substitutions per site. Spo.
= Sporisorium, Ust. = Ustilago
   () 135
Acknowledgements. We are grateful to Dr. Sándor Tóth (St. István University,
Gödöllő, Hungary) for preparing the Latin diagnosis, to Dr. Eric H.C.
McKenzie (Landcare Research, Auckland, New Zealand) and Dr. Michael Weiß
(University of Tübingen, Germany) for checking the manuscript , or part of it,
and for useful suggestions. We thank Dr. Michael Weiß, Dr. Sigisfredo Garnica
and Dr. Robert Bauer (University of Tübingen) for providing the molecular lab.
e technical assistance of Mrs. Christine Vánky with the illustrations (H.U.V.,
Tübingen, Germany), Mrs. Magda Eha-Wagner with preparation of semi-thin
sections, and Mrs. Monika Meinert (both from University of Tübingen) with
preparation of the SEM pictures of the spores, is gratefully acknowledged.
References
Altschul, S.F., Madden, T.L., Schä er, A.A., Zhang, J., Zhang, Z., Miller, W.
& Lipman, D.J. 1997. Gapped BLAST and PSI-BLAST: a new generation
of protein database search programs. — Nucleic Acids Research 25:
3389–3402.
Atkinson, G.F. 1897. Some fungi from Alabama. — Cornell University Science
Bulletin 3: 1–50.
Castresana, J. 2000. Selection of conserved blocks from multiple alignments
for their use in phylogenetic analysis. — Molecular Biology and Evolution
17: 540–552.
Clayton, W.D. & Renvoize, S.A. 1986. Genera graminum. Grasses of the
world. Kew Bulletin Additional Series 13. Royal Botanic Gardens, Kew,
London.
Crous, P.W., Gams, W., Stalpers, J.A., Robert, V. & Stegehuis, G. 2004.
MycoBank: an online initiative to launch mycology into the 21 century.
— Studies in Mycology 50: 19–22.
Felsenstein, J. 1981. Evolutionary trees from DNA sequences: a maximum
likelihood approach. — Journal of Molecular Evolution 17: 368–376.
Gardes, M. & Bruns, T.D. 1993. ITS primers with enhanced speci city for
basidiomycetes – application to the identi cation of mycorrhizae and rusts.
— Molecular Ecology 2: 113–118.
Gatesy, J., DeSalle, R. & Wheeler, W. 1993. Alignment-ambiguous nucleotide
sites and the exclusion of systematic data. — Molecular Phylogenetics and
Evolution 2: 152–157.
Giribet, G. & Wheeler, W.C. 1999. On gaps. — Molecular Phylogenetics
and Evolution 13: 132–143.
Huelsenbeck, J.P. & Ronquist, F. 2001. MRBAYES: bayesian inference of
phylogenetic trees. — Bioinformatics Applications Note 17: 754–755.
Huelsenbeck, J.P. & Rannala, B. 2004. Frequentist properties of Bayesian
posterior probabilities of phylogenetic trees under simple and complex
substitution models. — Systematic Biology 53: 904–913.
Huelsenbeck, J.P., Larget, B., Miller, R.E. & Ronquist, F. 2002. Potential
applications and pitfalls of Bayesian inference of phylogeny. — Systematic
Biology 51: 673–688.
Katoh, K. & Toh, H. 2008. Recent developments in the MAFFT multiple
sequence alignment program (outlines version 6). — Briefings in
Bioinformatics 9: 286–298.
Katoh, K., Misawa, K., Kuma, K. & Miyata, T. 2002. MAFFT: a novel method
for rapid multiple sequence alignment based on fast Fourier transform. —
Nucleic Acids Research 30: 3059–3066.
Katoh, K., Kuma, K., Toh, H. & Miyata, T. 2005. MAFFT version 5:
improvement in accuracy of multiple sequence alignment. — Nucleic Acids
Research 33: 511–518.
Ling, L. 1949. Taxonomic notes on Asiatic smuts. I. — Sydowia 3: 123–134.
Lutz, M., Bauer, R., Begerow, D., Oberwinkler, F. & Triebel, D. 2004.
Tuberculina, rust relatives attack rusts. — Mycologia 96: 614–626.
O’Donnell, K.L. 1993. Fusarium and its near relatives. — In: D.R. Reynolds &
J.W. Taylor [eds].  e fungal holomorph: mitotic, meiotic and pleomorphic
speciation in fungal systematic, pp. 225–233. CAB International,
Wallingford, UK.
Piepenbring, M. 2001. New species of smut fungi from the neotropics. —
Mycological Research 105[2000]: 757–767.
Piepenbring, M., Bauer, R. & Oberwinkler, F. 1998. Teliospores of smut fungi.
Teliospore connections, appendages, and germ pores studied by electron
microscopy; phylogenetic discussion of characteristics of teliospores. —
Protoplasma 204: 202–218.
Ronquist, F.R. & Huelsenbeck, J.P. 2003. MRBAYES 3: Bayesian phylogenetic
inference under mixed models. — Bioinformatics 19: 1572–1574.
Shivas, R.G., Vánky, K., Vánky, C., Kula, G.R. & Gavali, V. 2001. An
annotated check list of Ustilaginomycetes in Papua New Guinea. —
Australasian Plant Pathology 30: 231–237.
Silvestro, D. & Michalak, I. 2010. raxmlGUI: a graphical front-end for
RAxML. Available at http://sourceforge.net/projects/raxmlgui/.
Stamatakis, A. 2006. RAxML-VI-HPC: maximum likelihood-based
phylogenetic analyses with thousands of taxa and mixed models. —
Bioinformatics 22: 2688–2690.
Stamatakis, A., Hoover, P. & Rougemont, J. 2008. A rapid bootstrap algorithm
for the RAxML web servers. — Systematic Biology 57: 758–771.
Stoll, M., Begerow, D. & Oberwinkler, F. 2005. Molecular phylogeny of
Ustilago and Sporisorium, and related taxa based on combined analysis of
rDNA sequences. — Mycological Research 109: 342–356.
Sydow, H. & Petrak, F. 1928. Micromycetes philippinenses. Series prima. —
Annales Mycologici 26: 414–446.
irumalachar, M.J., Whitehead, M.D. & O’Brien, M.J. 1967. Studies in the
genus Tolyposporella. — Mycologia 59: 389–396.
Vánky, K. 1995. Taxonomical studies on Ustilaginales. XIII. — Mycotaxon
56: 197–216.
Vánky, K. 1998. Taxonomical studies on Ustilaginales. XVIII. — Mycotaxon
69: 93–115.
Vánky, K. 2002. Illustrated genera of smut fungi. 2 edn. APS Press, St.
Paul, Minnesota, USA.
Vánky, K. 2012. Smut fungi of the world. APS Press, St. Paul, Minnesota, USA.
Vánky, K. & Shivas, R.G. 2001. New smut fungi (Ustilaginomycetes) from
Australia. — Fungal Diversity 7: 145–174.
Vánky, K. & Shivas, R.G. 2008. Fungi of Australia.  e smut fungi. ABRS,
Canberra, CSIRO Publishing, Melbourne.
Vánky, K., Lutz, M. & Shivas, R.G. 2006. Anomalomyces panici, new genus
and species of Ustilaginomycetes from Australia. — Mycologia Balcanica
3: 119–126.
White, T.J., Bruns, T.D., Lee, S. & Taylor, J.W. 1990. Ampli cation and direct
sequencing of fungal ribosomal RNA genes for phylogenetics. — In: M.A.
Innis, D.H. Gelfand, J.J. Sninsky & T.J. White [eds]. PCR protocols, a guide
to methods and applications, pp. 315–322. Academic Press, San Diego.
Zundel, G.L. 1953.  e Ustilaginales of the world. — Pennsylvania State
College School of Agriculture Department of Botany Contribution 176:
I–XI + 1–410.
... the separation of Microbotryales from Ustilaginomycotina (Begerow et al. 1997(Begerow et al. , 2014, and division of the Ustilago-Sporisorium-Macalpinomyces complex into smaller, welldefined genera (McTaggart et al. 2012c). In the latter example, smut fungi on grasses in the Ustilago-Sporisorium-Macalpinomyces complex were divided into the genera Anthracocystis, Langdonia, Stollia, Triodiomyces and Tubisorus (Vánky & Lutz 2011, McTaggart et al. 2012c. ...
... Systematic studies showed that U. maydis was not closely related to species of Ustilago s. str., and was instead recovered as sister to species of Sporisorium and Anthracocystis (Piepenbring et al. 2002, Stoll et al. 2005, Vánky & Lutz 2011, McTaggart et al. 2012a. In these studies, U. maydis was closely related to U. bouriquetii, a smut fungus that forms hypertrophied sori in the inflorescences of Mycosarcoma is the earliest available generic name for the clade containing U. maydis, which was described as the type species (Brefeld 1912). ...
... Mycosarcoma is resurrected here and the circumscription emended to accommodate a monophyletic group in Ustilaginaceae; this addresses one further component of polyphyly in Ustilago s. lat. This taxonomy is supported by several separate systematic analyses that have determined a unique phylogenetic position of M. maydis within the family (Piepenbring et al. 2002, Stoll et al. 2005, Vánky & Lutz 2011, McTaggart et al. 2012a). We will submit a proposal to the Nomenclature Committee for Fungi for conservation of Uredo maydis over the name Lycoperdon zeae, which has priority at species rank, to avoid a disadvantageous nomenclatural change, as 'maydis' is an accepted and widely used epithet for corn smut in plant pathology and genetics. ...
Article
Full-text available
Ustilago is a polyphyletic genus of smut fungi found mainly on Poaceae. The development of a taxonomy that reflects phylogeny requires subdivision of Ustilago into smaller monophyletic genera. Several separate systematic analyses have determined that Macalpinomyces mackinlayi, M. tubiformis, Tolyposporella pachycarpa, Ustilago bouriquetii and U. maydis, occupy a unique phylogenetic position within the Ustilaginaceae. A previously introduced monotypic generic name typified by U. maydis, Mycosarcoma, is available to accommodate these species, which resolves one component of polyphyly for Ustilagos.lat. in Ustilaginaceae. An emended description of Mycosarcoma is provided to reflect the morphological synapomorphies of this monophyletic group. A specimen of Ustilago maydis that has had its genome sequenced is designated as a neotype for this species. Taxonomic stability will further be provided by a forthcoming proposal to conserve the name Uredo maydis over Lycoperdon zeae, which has priority by date, in order to preserve the well-known epithet maydis.
... Molecular phylogenetic studies by Stoll et al. (2003Stoll et al. ( , 2005 revealed that Sporisorium species are split into two main lineages. This finding was later confirmed by other molecular analyses (Cunnington et al. 2005;Vánky et al. 2006;Vánky and Lutz 2011;McTaggart et al. 2012a;Shivas et al. 2013;Zhang et al. 2013). These two main lineages included the type species of Sporisorium (S. sorghi Ehrenb. ...
... This study provides the most comprehensive phylogeny of Anthracocystis to date. It is the first phylogeny focused specifically on the genus Anthracocystis, as all previous phylogenetic studies either treated Anthracocystis as a synonym of Sporisorium in a wider context of grass-infecting Ustilaginales (Cunnington et al. 2005;Stoll et al. 2003Stoll et al. , 2005Vánky et al. 2006;Vánky and Lutz 2011;Zhang et al. 2013), or did not show relationships between species (McTaggart et al. 2012a, c;Shivas et al. 2013). The concatenated ITS+LSU dataset was constructed using only sequences that were linked to reliably identified (by smut experts) voucher specimens. ...
... Sporisorium scitamineum EF185083 was probably wrongly identified, as genuine Sporisorium scitamineum is placed in Sporisorium s. str. (Stoll et al. 2005;Vánky et al. 2006;Vánky and Lutz 2011). Sporisorium nervosum AY740057/ AY740110 generated from specimen M 56622 (Stoll et al. 2005) may also be wrongly identified, as a sequence of this species (HQ013106), generated from the holotype specimen (BRIP 27019), was nested within the Sporisorium s. str. ...
Article
Full-text available
The genus Anthracocystis (Ustilaginales, Ustilaginaceae) was recently reinstated for grass-infecting species of smut fungi that have sori with a peridium composed of mostly fungal cells, filiform or slender columellae, persistent spore balls usually composed of dimorphic spores, and lacking sterile cells between spore balls. In this study, Anthracocystis grodzinskae sp. nov. on Euclasta condylotricha is described and illustrated from the Sudanian savanna biome in Benin (West Africa). The new species is compared with two other smut fungi known on Euclasta condylotricha, namely Sporisorium euclastae and Anthracocystis ischaemoides, in Zambia. It differs from these species in a number of morphological characters that are discussed in detail . The systematic position of A. grodzinskae was investigated in a phylogenetic analysis with a concatenated supermatrix of the internal transcribed spacer (ITS) and large subunit (LSU) regions of ribosomal DNA. The dataset included all representatives of Anthracocystis for which sequences were available in the National Center for Biotechnology Information's (NCBI’s) GenBank and that were linked to reliably identified source specimens, related yeast species, and unnamed yeast strains or environmental sequences. The phylogenetic hypothesis derived from the dataset is intended to serve as a backbone tree for Anthracocystis. 19 ITS and 13 LSU sequences were tracked to represent sequences generated from type specimens (holotypes, isotypes or paratypes). These type sequences are recommended to be deposited in the RefSeq Targeted Loci database. This study provides the first explicit evidence that several asexual species are nested within the Anthracocystis lineage. The yeast sequences were scattered in different subclades of Anthracocystis and none of them could be directly assigned to a teleomorphic species. Only one of these yeast anamorphs was assigned to a species, namely Pseudozyma flocculosa. In line with the current code of nomenclature, and following recent practice of merging yeast species with sexual species under the older generic name, this yeast is recombined into Anthracocystis as A. flocculosa. Additionally, new combinations are proposed for four teliosporic species (Anthracocystis andrewmitchellii, A. christineae, A. kenyana, A. warambiensis).
... It consists of 14 genera, including Ustilago, Sporisorium, and Macalpinomyces [1]. The taxonomy of the Ustilaginaceae, which is still a matter of debate, was focused on during several studies resulting in several taxonomic revisions [2]. For instance, Kirk et al. (2008) proposed that the Ustilaginaceae family comprises 17 genera comprising 607 species [3]. ...
Article
Full-text available
The family of Ustilaginaceae belongs to the order of Basidiomycetes. Despite their plant pathogenicity causing, e.g., corn smut disease, they are also known as natural producers of value-added chemicals such as extracellular glycolipids, organic acids, and polyols. Here, we present 17 high-quality draft genome sequences (N50 > 1 Mb) combining third-generation nanopore and second-generation Illumina sequencing. The data were analyzed with taxonomical genome-based bioinformatics methods such as Percentage of Conserved Proteins (POCP), Average Nucleotide Identity (ANI), and Average Amino Acid Identity (AAI) analyses indicating that a reclassification of the Ustilaginaceae family might be required. Further, conserved core genes were determined to calculate a phylogenomic core genome tree of the Ustilaginaceae that also supported the results of the other phylogenomic analysis. In addition, to genomic comparisons, secondary metabolite clusters (e.g., itaconic acid, mannosylerythritol lipids, and ustilagic acid) of biotechnological interest were analyzed, whereas the sheer number of clusters did not differ much between species.
... Cereal Pathogens: fungal pathogenic spores Pericornia and Sorosporium (Kálmán and Matthias, 2011;Gramaje et al., 2020). ...
Article
Long-standing arguments regarding the early cultural transition, the domestication of plants and the impacts of climate change on past Egyptian societies remain contentious. In this paper, we demonstrate that grazing started at our study site, Kom El-Khilgan in the NE Nile Delta, ca. 7000 years ago, which was several hundred years prior to crop farming. We examined pollen-spores and non-pollen palynomorphs (NPP) in a 1-m-deep sediment profile (KH-1) in the study site, defined archaeologically as the Pre-Dynastic (>5.9 ka) in age. Our results show that before ca. 7.0 ka, major floods prevailed in the Nile Delta, as highly-concentrated Podocarpus and Polypodiaceae were transported from the East African highlands and Cyperaceae of higher Nile flow indication. This humid phase was succeeded by a brief period of drying climate ca. 7.0–6.6 ka, allowing the entry of the first settlers who commenced grazing their animals on previously inundated wetlands. This change is indicated by the remarkable increase in fungal spores (Cercophora, Sordaria, Coniochaeta cf. Ligniaria) from accumulations of animal dung. At ca. 6.6 ka, the abrupt appearance of domesticated cereal pollen (Poaceae >35 μm) and cereal grass pathogens (Pericornia and Sorosporium) suggests an amelioration of the climate that allowed the introduction of cereal crops and related water management activities. Grazing and cropping co-existed for the remainder of the record during the time when there occurred a mega-tendency of climate drying towards recent time. A drought event recognized ca. 4.2 ka led to the collapse of the Old Kingdom.
... Rabenh., and Ustanciosporium standleyanum (Zundel) M. Piepenbr., parts of the concatenated sequences of the remaining species were obtained from different specimens/cultures of the species (for detailed information, compare GenBank accessions). For GenBank accession numbers of the sequences of dataset 1 (Boekhout et al. 1995(Boekhout et al. , 2003Begerow et al. 1997Begerow et al. , 2000Begerow et al. , 2001Begerow et al. , 2002aBegerow et al. , 2006Bauer et al. 1999Bauer et al. , 2001Bauer et al. , 2005Bauer et al. , 2007Bauer et al. , 2008Piepenbring et al. 1999Piepenbring et al. , 2002Piepenbring et al. , 2010Fell et al. 2000;Castlebury et al. 2005;Hendrichs et al. 2005;Stoll et al. 2005;Maier et al. 2006;Matheny et al. 2006;Vánky et al. 2006Vánky et al. , 2008Vánky et al. , 2013González et al. 2007;Chandra and Huff 2008;Lutz et al. 2008Lutz et al. , 2012aPaap et al. 2008;Ritschel et al. 2008;Tanaka et al. 2008;Sipiczki and Kajdacsi 2009;Deadman et al. 2011;Vánky and Lutz 2011;McTaggart et al. 2012;Piątek et al. 2013;Nasr et al. 2014), see Fig. 3. For GenBank accession numbers of the sequences of dataset 2 (de Wachter et al. 1992;Boekhout et al. 1995;Schillberg et al. 1995;Swann and Taylor 1995;Takashima and Nakase 1996;Bakkeren et al. 2000;Döring and Blanz 2000;Hamamoto et al. 2000;Döring 2003;Liu and Hall 2004;Lutzoni et al. 2004;Wingfield et al. 2004;Castlebury et al. 2005;Stoll et al. 2005;Begerow et al. 2006;de Beer et al. 2006;Kolařík et al. 2006;Matheny et al. 2006Matheny et al. , 2007Carris et al. 2007;Le Gac et al. 2007;Ran et al. 2008;Brock et al. 2009;Rosa et al. 2009;Kottke et al. 2010;Gorfer et al. 2011;Schoch et al. 2012;Wang et al. 2014), see Table 2. ...
Article
Full-text available
The order Ceraceosorales (Ustilaginomycotina) currently includes the single genus Ceraceosorus, with one species, Ceraceosorus bombacis, parasitic on Bombax ceiba in India. The diversity, biogeography, evolution, and phylogenetic relationships of this order are still relatively unknown. Here, a second species of Ceraceosorus is described from West Africa as a novel species, Ceraceosorus africanus, infecting Bombax costatum in Benin, Ghana, and Togo. This species produces conspicuous fructifications, similar to corticioid basidiomata when mature, but sorus-like in early stages of ontogenetic development. The fructifications cover much of the leaf surface and resemble leaf blight. This contrasts with the inconspicuous fructifications of C. bombacis comprising small spots scattered over the lower leaf surface that resemble leaf spot. Both species of Ceraceosorus differ in several micromorphological traits, infect different host plant species in widely separated geographical areas, and are separated by a considerable genetic distance in 28S rDNA and RPB2 genes. The distinct corticioid fructification of C. africanus is a unique morphological trait within the Ustilaginomycotina. Molecular phylogenetic analyses of a single gene dataset (D1/D2 28S rDNA) supported the monophyly of the two Ceraceosorus species and the Ceraceosorales and their placement within the Ustilaginomycotina. Molecular phylogenetic analyses of a multigene dataset (18S/5.8S/28S rDNA/RPB2/TEF1) revealed Exobasidium rhododendri (Exobasidiales) as the closest relative of Ceraceosorus, both clustering together with Entyloma calendulae (Entylomatales), indicating affinities to the Exobasidiomycetes. This phylogenetic placement is in agreement with ultrastructural characteristics (presence of local interaction zone and interaction apparatus) reported for the Ceraceosorales, Entylomatales, and Exobasidiales.
... Confirmation of specimen identity is enhanced by comparison with high quality images of authentic reference specimens taken in the field and in the laboratory. Since the last revision of smut fungi in Australia (), six new generic names with type species from Australia have been established: Aizoago (), Langdonia (McTaggart et al. 2012b), Shivasia (Lutz et al. 2012), Stollia (McTaggart et al. 2012b), Triodiomyces (McTaggart et al. 2012b), and Tubisorus (Vánky & Lutz 2011). The genus Anthracosystis was resurrected (McTaggart et al. 2012b), and the first representative of Aurantiosporium from Australia was collected (Table 1). ...
Article
Full-text available
Interactive identification keys for Australian smut fungi (Ustilaginomycotina and Pucciniomycotina, Microbotryales) and rust fungi (Pucciniomycotina, Pucciniales) are available online at http://collections.daff.qld.gov.au. The keys were built using Lucid software, and facilitate the identification of all known Australian smut fungi (317 species in 37 genera) and 100 rust fungi (from approximately 360 species in 37 genera). The smut and rust keys are illustrated with over 1,600 and 570 images respectively. The keys are designed to assist a wide range of end-users including mycologists, plant health diagnosticians, biosecurity scientists, plant pathologists, and university students. The keys are dynamic and will be regularly updated to include taxonomic changes and incorporate new detections, taxa, distributions and images. Researchers working with Australian smut and rust fungi are encouraged to participate in the ongoing development and improvement of these keys.
... Ustilago, Sporisorium and Macalpinomyces are the three genera of smut fungi (subphylum Ustilaginomycotina). These three genera are comprised of approximately 530 described species that all infect grasses [3]. Compared to the ambiguous position of Macalpinomyces, previous studies have demonstrated that Ustilago and Sporisorium together form a monophyletic group with the Ustilaginomycotina [4,5]. ...
Article
Full-text available
(URL+Background+Results+Conclusions) URL: http://www.biomedcentral.com/1471-2164/15/996 Background : Sugarcane smut can cause losses in cane yield and sugar content that range from 30 % to total crop failure. Losses tend to increase with the passage of years. Sporisorium scitamineum is the fungus that causes sugarcane smut. This fungus has the potential to infect all sugarcane species unless a species is resistant to biotrophic fungal pathogens. However, it remains unclear how the fungus breaks through the cell walls of sugarcane and causes the formation of black or gray whip-like structures on the sugarcane plants. Results : Here, we report the first high-quality genome sequence of S. scitamineum assembled de novo with a contig N50 of 41 kb, a scaffold N50 of 884 kb and genome size 19.8 Mb, containing an estimated 6,636 genes. This phytopathogen can utilize a wide range of carbon and nitrogen sources. A reduced set of genes encoding plant cell wall hydrolytic enzymes leads to its biotrophic lifestyle, in which damage to the host should be minimized. As a bipolar mating fungus, a and b loci are linked and the mating-type locus segregates as a single locus. The S.scitamineum genome has only 6 G protein-coupled receptors (GPCRs) grouped into five classes, which are responsible for transducing extracellular signals into intracellular responses, however, the genome is without any PTH11-like GPCR. There are 192 virulence associated genes in the genome of S. scitamineum, among which 31 expressed in all the stages, which mainly encode for energy metabolism and redox of short-chain compound related enzymes. Sixty-eight candidates for secreted effector proteins (CSEPs) were found in the genome of S. scitamineum, and 32 of them expressed in the different stages of sugarcane infection, which are probably involved in infection and/or triggering defense responses. There are two non-ribosomal peptide synthetase (NRPS) gene clusters that are involved in the generation of ferrichrome and ferrichrome A, while the terpenes gene cluster is composed of three unknown function genes and seven biosynthesis related genes. Conclusions : As a destructive pathogen to sugar industry, the S. scitamineum genome will facilitate future research on the genomic basis and the pathogenic mechanisms of sugarcane smut.
... Additionally, the sequence of an uncultured fungus clone, MAC09GadID (GenBank: JN890075), revealed by a Blast search (Altschul et al. 1997) as a closely related sequence, was added to the LSU dataset. GenBank accession numbers of the sequences used for the LSU dataset (Boekhout et al. 1995;Begerow et al. 1997Begerow et al. , 2007Piepenbring et al. 1999Piepenbring et al. , 2002Fell et al. 2000;Castlebury et al. 2005;Hendrichs et al. 2005;Stoll et al. 2005;Vánky et al. 2006Vánky et al. , 2008aBauer et al. 2007Bauer et al. , 2008González et al. 2007;Matheny et al. 2007;McGuire et al. 2010;Vánky and Lutz 2011;Lutz et al. 2012a, b;McTaggart et al. 2012;Piątek et al. 2013) are given in Fig. 1. ...
Article
Full-text available
The systematic position of two strains of a yeast-like fungus isolated from plant remnants on the Kharg Island in the Persian Gulf of Iran is evaluated using morphological, physiological and phylogenetic analyses. In culture, this fungus produced cylindrical cells that reproduced by polar budding on short stalks. Production of ballistoconidia and blastospores was observed. The carbon source assimilation spectrum was broad, but fermentation ability was absent. Phylogenetic analyses of the nuclear SSU, LSU (D1/D2 domain), and ITS rDNA revealed that this fungus represents a new lineage in the Urocystidales of the subphylum Ustilaginomycotina. Based on the comparison of phenotypic characters, physiology, and DNA sequences, a new genus and species Fereydounia khargensis (IBRC-M 30116T = CBS 13305T) is described for this fungus and accommodated in the novel family Fereydouniaceae. This is the first report of anamorphic saprobic fungus residing in the Urocystidales, stressing the remarkable evolutionary diversity in the subphylum Ustilaginomycotina.
Article
Full-text available
Fungi are an important and diverse component in various ecosystems. The methods to identify different fungi are an important step in any mycological study. Classical methods of fungal identification, which rely mainly on morphological characteristics and modern use of DNA based molecular techniques, have proven to be very helpful to explore their taxonomic identity. In the present compilation, we provide detailed information on estimates of fungi provided by different mycologistsover time. Along with this, a comprehensive analysis of the importance of classical and molecular methods is also presented. In orderto understand the utility of genus and species specific markers in fungal identification, a polyphasic approach to investigate various fungi is also presented in this paper. An account of the study of various fungi based on culture-based and cultureindependent methods is also provided here to understand the development and significance of both approaches. The available information on classical and modern methods compiled in this study revealed that the DNA based molecular studies are still scant, and more studies are required to achieve the accurate estimation of fungi present on earth.
Article
Full-text available
Many fungi are pathogenic on plants and cause significant damage in agriculture and forestry. They are also part of the natural ecosystem and may play a role in regulating plant numbers/density. Morphological identification and analysis of plant pathogenic fungi, while important, is often hampered by the scarcity of discriminatory taxonomic characters and the endophytic or inconspicuous nature of these fungi. Molecular (DNA sequence) data for plant pathogenic fungi have emerged as key information for diagnostic and classification studies, although hampered in part by non-standard laboratory practices and analytical methods. To facilitate current and future research, this study provides phylogenetic synopses for 25 groups of plant pathogenic fungi in the Ascomycota, Basidiomycota, Mucormycotina (Fungi), and Oomycota, using recent molecular data, up-to-date names, and the latest taxonomic insights. Lineage-specific laboratory protocols together with advice on their application, as well as general observations, are also provided. We hope to maintain updated backbone trees of these fungal lineages over time and to publish them jointly as new data emerge. Researchers of plant pathogenic fungi not covered by the present study are invited to join this future effort. Bipolaris, Botryosphaeriaceae, Botryosphaeria, Botrytis, Choanephora, Colletotrichum, Curvularia, Diaporthe, Diplodia, Dothiorella, Fusarium, Gilbertella, Lasiodiplodia, Mucor, Neofusicoccum, Pestalotiopsis, Phyllosticta, Phytophthora, Puccinia, Pyrenophora, Pythium, Rhizopus, Stagonosporopsis, Ustilago and Verticillium are dealt with in this paper.
Article
Seven new species of smut fungi from the neotropics are described. The new species are remarkable for their special morphological development, as new representatives of small genera, or by interesting host plants and contribution to our knowledge of neotropic biodiversity. Aurantiosporium pallidum from Bolivia is the third species known in its genus. Its sori develop in the hypertrophied tissue of the rachillae of female and male spikelets. Kuntzeomyces ruizianae from Colombia is the second species known in its genus, characterized by relatively small teliospores. Moreaua bulbostylidis from Bolivia and Venezuela is the first record of a species of this genus (formerly Tolyposporium) on the genus Bulbostylis (Cyperaceae). Thecaphora amaranthicola from Ecuador differs from the known species T. amaranthi on the same host genus by smaller balls of teliospores. Thecaphora smallanthi is the first species of Thecaphora known to include hyaline globose sterile cells in the centre of teliospore balls. Tilletia boliviana is the first species of Tilletia infecting a species of Bromus (Poaceae) and forming warty teliospores. Species of Urocystis on Poaceae are known as agents of stripes on leaves, stems, and inflorescences. Urocystis reinhardii from Bolivia is the first species of this genus with sori in the interior of hypertrophic peduncles of a grass.
Article
A multiple sequence alignment program, MAFFT, has been developed. The CPU time is drastically reduced as compared with existing methods. MAFFT includes two novel techniques. (i) Homo logous regions are rapidly identified by the fast Fourier transform (FFT), in which an amino acid sequence is converted to a sequence composed of volume and polarity values of each amino acid residue. (ii) We propose a simplified scoring system that performs well for reducing CPU time and increasing the accuracy of alignments even for sequences having large insertions or extensions as well as distantly related sequences of similar length. Two different heuristics, the progressive method (FFT-NS-2) and the iterative refinement method (FFT-NS-i), are implemented in MAFFT. The performances of FFT-NS-2 and FFT-NS-i were compared with other methods by computer simulations and benchmark tests; the CPU time of FFT-NS-2 is drastically reduced as compared with CLUSTALW with comparable accuracy. FFT-NS-i is over 100 times faster than T-COFFEE, when the number of input sequences exceeds 60, without sacrificing the accuracy.