Você está na página 1de 14

Fungal Genetics and Biology 56 (2013) 1-146

Contents lists available at SciVerse ScienceDirect

Fungal Genetics and Biology


ELSEVIER journal homepage: www.elsevier.com/locate/yfgbi

The bacterial secondary metabolite 2,4-diacetylphloroglucinol impairs CrossMark

mitochondrial function and affects calcium homeostasis in Neurospora


crassa

Danielle M. Troppensa, Meiling Chuc, LucyJ. Holcombea, Olive Gleesona, Fergal OGarab, Nick D. Readc 1, John P.
Morrissey^*
Microbiology Department, University College Cork, Cork, Ireland ab Biomerit Research Centre, Biosciences Institute, University College Cork, Cork, Ireland
c
Fungal Cell Biology Group, Institute of Cell Biology, University ofEdinburgh, Rutherford Building, Edinburgh EH9 3JH, UK

A R T I C L E I N F Q

Article history:
ived 6 February 2013 Accepted A B S T R A C T
12 April 2013 Available online T h e bacterial secondary metabolite 2,4-diacetylphloroglucinol (DAPG) is of interest as an active ingredi- ent of
23 April 2013 biological control strains of Pseudomonas fluorescens and as a potential lead pharmaceutical mol- ecule because of its capacity to inhibit
growth of diverse microbial and non-microbial cells. The mechanism by which this occurs is unknown and in this study the filamentous
Keywords: fungus Neurospora crassa was used as a model to investigate the effects of DAPG on a eukaryotic cell. Colony growth, conidial ger- mination
G and cell fusion assays confirmed the inhibitory nature of DAPG towards N. crassa. A number of different fluorescent dyes and fluorescent
ospora crassa protein reporters were used to assess the effects of DAPG treat- ment on mitochondrial and other cellular functions. DAPG treatment led to
Mitochondria changes in mitochondrial morphology, and rapid loss of mitochondrial membrane potential. These effects are likely to be respon- sible for
Calcium the toxicity of DAPG. It was also found that DAPG treatment caused extracellular calcium to be taken up by conidial germlings leading to a
Antimicrobial transient increase in cytosolic free Ca2+ with a distinct con- centration dependent Ca 2+ signature.

2013 Elsevier Inc. All rights reserved


.
1. Introduction

Microbes have the capacity to synthesise a diverse array of secondary


metabolites such as phenolics, polyketides, cationic pep- tides, non-ribosomal
peptides, alkyl quinolones, oxylipins and homoserine lactones. Some of these
metabolites have clearly de- fined and validated functions. Siderophores, for
instance are se- creted to chelate iron for the producing microbe (Cornelis, 2010),
N-acyl homoserine lactones are used to signal within a microbial population (Ng
and Bassler, 2009; Williams and Camara, 2009), and pheromones are used to
induce mating within a species (Jones and Bennett, 2011). Many secondary
metabolites have also been reported to have antimicrobial properties and indeed,
most clinical antibiotics are derived from metabolites naturally produced by
bacteria or fungi (Demain, 2009; Yim et al., 2006). In addition, many more
secondary metabolites are capable of inhibiting micro- bial growth under
laboratory conditions. This has led to the general consensus that the biological
role of most secondary metabolites is to antagonise other microbes in the
environment to gain a compet- itive advantage. More recently, Davies and
colleagues have strongly challenged the prevailing wisdom on the ecological roleof
antibiotics (Yim et al., 2006, 2007). Their thesis, founded on tran- scriptome
analyses, is that many so-called antibiotics act as intra- or inter-microbial signal
molecules at concentrations far below their minimum inhibitory concentrations
(MICs) (Goh et al., 2002). Subsequent work from a number of groups has
provided some support for the notion that certain secondary metabolites with
antibiotic activity at high concentrations have signalling func- tions at lower

1 Corresponding authors.
E-mail address: m.j.morrissey@ucc.ie (J.P. Morrissey).

1087-1845/$ - see front matter 2013 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.fgb.2013.04.006
2 D.M. Troppens et al./Fungal Genetics and Biology 56 (2013) 135-146

concentrations, an effect termed hormesis (Cummins et al., 2009; olites, with species within the genera Streptomyces, Bacillus and
Duan et al., 2009; Kuroda et al., 2007; Shen et al., 2008; Yim et Pseudomonas best known in this regard (Challis and Hopwood,
al., 2006). Research in yeast indicates that hor- mesis may be a 2003; Gross and Loper, 2009; Raaijmakers et al., 2010; Stein,
common phenomenon. For example, a study of >2000 potential 2005). Whereas the anthropocentric focus with Streptomycetes
anti-cancer drugs established that many displayed a hormetic has been on the production of antibiotics with clinical potential,
effect (Calabrese et al., 2008) and recently it was pro- posed that many studies have sought to exploit the intrinsic activities of
hormesis effects of hydrogen peroxide could influence cellular specific Bacillus and Pseudomonas strains to develop bacterial
aging in yeast (Mesquita et al., 2010). inoculants that can be used in agriculture as natural biocontro
Many soil microbes are prolific producers of secondary metab-
l isolates, no clues to intracellular targets were revealed by those
studies (Schouten et al., 2008, 2004).
agents to control phytopathogenic fungi (Bloemberg and
Lugten- berg, 2001; Haas and Defago, 2005; Morrissey et al., In this study, Neurospora crassa was selected as a model to
2004; Raaij- makers et al., 2002). Pseudomonas species have as- sess the effects of DAPG on a filamentous fungus. N. crassa
attracted a lot of research as particular strains produce was the original fungal genetic model and in recent years has
metabolites, such as phena- zines, cyclic lipopeptides and been extensively used as a model organism for fungi and
polyketides that have antifungal activity in the laboratory and eukaryotes (Davis and Perkins, 2002). Many cell biology,
potential applications in biotechnol- ogy. One secondary microscopy and ge- netic tools are available with N. crassa and
metabolite of particular interest is 2,4-diac- etylphloroglucinol this fungus has proved useful to study the effects of small
(DAPG), which is produced mainly by the rhizosphere-associated molecules such as antifungal peptides and proteins (Binder et al.,
bacterium Pseudomonas fluorescens (Ban- gera and Thomashow, 2010; Munoz et al., 2013, 2012; Spelbrink et al., 2004; Thevissen
1996; Shanahan et al., 1993). This phenolic metabolite is et al., 1999), azoles (Selker, 1998), chitosan (Palma-Guerrero et
synthesised by the activities of the PhlD polyketide synthase and al., 2009, 2010), and other antifungal drugs or metabolites
PhlACB acetlytransferase enzymes (Achkar et al., 2005) and (Ermolayeva and Sanders, 1995; Pere- ira and Said, 2009; Selker,
possesses broad spectrum growth inhibitory activity to- wards 1998). In this study, we investigated the effects of DAPG on N.
phytopathogenic fungi (Keel et al., 1996; Kwak et al., 2009; Laville crassa to determine what processes and func- tions were
et al., 1992; Mazzola et al., 1995; Vincent et al., 1991), bacteria impaired or affected by this bacterial metabolite.
(Cronin et al., 1997b; Keel et al., 1992), oomycetes (de Souza et 2. Materials and methods
al., 2003; Fenton et al., 1992), and nematodes (Cronin et al.,
1997a). Because of this broad spectrum activity against plant 2.1. Strains, plasmids and culture conditions
patho- gens, many studies have sought to exploit the natural
plant growth/ health promoting activity of P. fluorescens to All N. crassa strains used in this study are listed in Table 1.
develop biocontrol inoculants (reviewed in: (Haas and Defago, To measure cytosolic free Ca2+ ([Ca2+]c) levels, either wild type (wt)
2005; Morrissey et al., 2002; Weller et al., 2007)). Interestingly, a containing the pAZ6 aeqS expression vector or mutant strains
number of studies have reported that DAPG can also trigger a con- taining the pAB19 aeqS expression vector were used (Binder
global defence response in plants known as induced systemic et al., 2010). All strains were cultured and maintained on Vogels
resistance (ISR) through a mech- anism yet to be determined med- ium containing 2% sucrose (Davis, 2000). To select for
(Iavicoli et al., 2003; Siddiqui and Shau- kat, 2004; Verhagen et mutant strains, medium was supplemented with hygromycin. To
al., 2010; Weller et al., 2007). select for aeqS expressing strains, medium was supplemented
with hygromycin for the wild type or ignite (glufosinate) for
The diverse effects of DAPG on microbes and plants raise mutant strains, respectively. Agar slants and plates were
inter- esting questions as to its targets, mode-of-action and incubated at either 25 C or 35 C with or without light (as
ecological function. At one level, it can be considered a classical indicated). To promote conidiation, plates and slants were
antimicrobial, whereas the finding that most phlD+ strains incubated at room tem- perature with continuous light.
produce only low lev- els of DAPG, and the intriguing effects on
plants, may be more con- sistent with a signalling role. Some 2.2. Chemicals and dyes
studies have addressed the question of molecular effects of DAPG
on susceptible organisms. Exposure of the oomycete Pythium DAPG was synthesised in-house as previously described and
ultimum to DAPG led to inhibi- tion of hyphal growth and stored as a powder (Shanahan et al., 1993). A stock of 10 mg
zoospore motility as well as structural changes to the cell DAPG ml-1 in methanol was stored at -20 C. 1,2-Bis(o-amino-
including alteration of the plasma membrane, vacuolization and phenoxy)ethane-N,N,N',N'-tetraacetic acid (BAPTA) tetrapotassium
cell content disintegration (de Souza et al., 2003). A study of a salt (Sigma) was dissolved in distilled water, antimycin A (Cam-
library of Saccharomyces cerevisiae (yeast) mu- tants identified bridge Biosciences) stock solution was dissolved in 96% ethanol
multiple mutants in different cellular processes that were more and carbonylcyanide-3-chlorophenylhydrazone (CCCP) (Sigma)
sensitive to DAPG than the wild-type strain (Kwak et al., 2010). stock solution was dissolved in DMSO. Methylsalicylate and coel-
These data support the idea that DAPG may affect fundamental enterazine stock solutions were dissolved in methanol.
cellular processes, a suggestion also made in other stud- ies Fluorescent dyes used in this study are listed in Table 2.
(Jousset, 2006). Direct studies of physiological functions in yeast
demonstrated that DAPG treatment led to loss of mitochon- drial 2.3. Colony extension assay
function, indicating that DAPG may act to depolarise mito-
chondrial membranes (Gleeson et al., 2010). Studying defence
The measurement of the radial extension of a colony was
mechanisms can sometimes yield insights into the targets of an done as previously described (Berepiki et al., 2010), with slight
antimicrobial compound but although studies of the fungi
modifica- tion. A conidial suspension was prepared in sterile
Botrytis cinerea and Fusarium oxysporum implicated degradative distilled water and filtered, and a 10 pl conidial suspension was
enzymes and ABC transporters in tolerance in some natural placed onto the centre of an agar plate containing no
D.M. Troppens et al./Fungal Genetics and Biology 56 (2013) 135-146 3

supplementation, DAPG at concentrations ranging from 2.5 to 15 were counted using a haemocytometer and diluted to 5 x 105-
pgml-1, or 2% methanol (the DAPG solvent). Four radii were conidia/ml in Vogels medium which was transferred to an 8-well
randomly marked on the underside of each plate. The plates were culture chamber (Nalg Nunc, Rochester, NY) with 200 pl per well.
then incubated at 35 C in the dark and colony extension was DAPG (dissolved in methanol) at the appropriate final concentra-
measured after 28 h of growth. Colony extension was expressed tion, or equal volumes of methanol or Vogels medium, was added
as percentage of average radius on treatment plates (methanol or to the wells. Chambers were transferred to 25 C in the dark and
DAPG) compared to control plates (no supplement). The incubated for 6 h. At least 10 random areas containing conidia
experiment was done at least in duplicate per concentration and and/or germlings were viewed per treatment using a 60 x
strain. objective on an inverted Eclipse TE2000E microscope with a
DXM1200F camera and ACT-1 image capture software (Nikon,
2.4. Germination and CAT fusion assay Kingston- Upon-Thames, Surrey, UK). Germination was expressed
as the per- centage of germinated conidia possessing germ tubes
A conidial suspension was prepared in Vogels medium from 4 to and/or conidial anastomosis tubes (CATs) compared to the total
5-day old cultures incubated at 35 C with light. Filtered conidia amount of conidia/germlings (n > 100). CAT fusion was expressed
as th
e
4 D.M. Troppens et al./Fungal Genetics and Biology 56 (2013) 135-146

Table 1
Strains used in this study.

Strain FGSC number Mating type Genotype Source

Table 2

a PI Sigma H O 1.5 mM 7.5 nM 550 625


Final Solutions diluted in Vogel's liquid mdium. b2Values in brackets are for confocal microscopy.

percentage of conidia or germlings involved in fusion events com-


pared to the total number of conidia/germlings (n > 100) (Roca et
al., 2005). This experiment was done in duplicate.

2.5. Measurement of intracellular Ca2+

A conidial suspension was prepared in Vogels medium from a


7-day old culture of wild-type or mutants strains (Dcch-1, Dfig-1,
Dyvc-1) expressing aequorin were incubated at 35 C with contin-
uous light. A conidial suspension was diluted to 10 6 conidia/ml
in Vogels liquid medium and coelenterazine was added to a final
concentration of 5 pM. 100 pl of conidial suspension was then
transferred to 6 wells of a 96-well microtitre plate per treatment.
The 96-well plates were incubated at 25 C in the dark for 6 h.
Each plate was then gently loaded into a LB96P Microlumat
lumino- meter or a Microlumat Plus LB96V plate luminometer
(Berthold Technology, Bad Wilbad, Germany). To measure
changes in cyto- solic free Ca2+ ([Ca2+]c), 100 pl DAPG at 2x the
final concentration ranging from 15 to 120 pgmr1 was
automatically injected into the wells (6 wells per treatment) and
luminescence was measured for a maximum of 11 min. To
measure the total luminescence of aequorin expressed by cells in
each well, 100 pl of 20% (vol/vol) ethanol containing 3 M CaCl 2
was injected into separate wells to allow total discharge of
aequorin by permeabilisation of the cells. To chelate extracellular
Ca2+, 100 pl BAPTA was added to one control row of wells 10 min
prior to DAPG injection at a final concentration of 7.5 mM. In that
case the injected DAPG was 3x final concentration. To monitor
luminescence and control injections of the plate reader, the PC
running Microsoft Windows based Bert- hold WinGlow software
was used. To convert luminescence values to [Ca2+]c
concentrations, the TermBert software developed by Zelter et al.
(2004) was used. Amplitudes of highest concentrations of DAPG
were tested for statistical difference by performing a t-test (n = 4)
using SigmaStat 3.5.

2.6. Fluorescence microscopy


D.M. Troppens et al./Fungal Genetics and Biology 56 (2013) 135-146 5

A conidial suspension was prepared in Vogels medium from a 3 (final concentration of 7.5 mM) was then added and cells
to 4-day old culture incubated at 35 C in the light. The conidial incubated for 10 min before addition of dye and DAPG. To assess
suspension was diluted to 5 x 105 conidia/ml in Vogels medium the effects of a longer exposure to DAPG, conidia were incubated
and transferred to an 8-well slide culture chamber (Nalg Nunc) for 3 h at 25 C, DAPG was added at a final concentration of 60
with 180 pl per well and grown for 5 h at 25 C. The slide culture pg ml-1 and the conidia were incubated for a further 2 h before
chamber was then transferred to a Nikon Eclipse TE 2000E the addition of dye and imaging as described above. An Eclipse
inverted microscope and a fluorescent dye (DASPMI, rhodamine TE 2000E inverted microscope (Nikon) was used for widefield light
123, or rho- damine B; Table 2) added to enable imaging of microscopy. Fluorescence images were captured with an EM-CCD
germlings. Where required, chemical treatments were added at an Orca camera (Hamamatsu) and Image Pro and MetaMorph
appropriate con- centration to bring the final volume in the wells software (Molecular Devices, Sunnydale, CA), respec- tively, and
to 200 pl and the sample was immediately viewed by fluorescence differential interference contrast (DIC) images were captured with
microscopy. Images of the same region were captured every a DXM1200F camera and ACT-1 image capture software (Nikon).
minute for a max- imum of 10 min (following DAPG or CCCP For confocal microscopy, an inverted TE 2000U Eclipse
treatment) and 20 min (following antimycin A treatment), microscope (Nikon) coupled to a Radiance 2100 confocal system
respectively, using a 100x objective. The final concentrations of with Lasersharp 2000 software v. 5.1 (Bio-Rad Micro- science,
DAPG used ranged from 15 to 60 pgmr1, CCCP was added at a Hemel Hempsted, UK) were used. An H1-GFP expressing strain
final concentration of 2 pM and antimycin A was added at a final (Table 1), in which nuclei were labelled with GFP, was trea- ted
concentration of 0.5 pg ml-1. Equivalent volumes of methanol or with DAPG for 2 h as described above and stained with rhoda-
Vogels medium were added as controls. For pre-treatment with mine B. Final fluorescent dye concentrations and the microscope
BAPTA, the conidial suspension was diluted to 106 conidia/ml settings used for fluorescent dyes and GFP are listed in Table 2.
with 80 pl added to each well and grown for 5 h. 100 pl of BAPTA To quantify permeabilisation by DAPG, a conidial suspension wa
s

stained with propidium iodide (PI) and fluorescein diacetate (FDA), treated with 15 or 60 pg DAPG mi -1 and imaged immediately. As
a control 70% ethanol was added. The percentage of permeabilised conidia was calculated (n > 100).

3. Results

3.1. DAPG inhibits colony growth and germination and CAT fusion

To determine the effects of the bacterial metabolite DAPG on the fungus N. crassa, assays were carried out to assess colony
extension rates, conidial germination and CAT fusion in the pres- ence of increasing concentrations of DAPG (Fig. 1). Colony extension
on agar plates containing Vogels medium was inhibited by levels as low as 2.5 pg DAPG ml -1 (65% reduction in colony diam- eter), with
a progressively greater inhibition as the DAPG concentraron increased (Fig. 1A). Treatment of germinating conidia with DAPG also
resulted in dose-dependent inhibition of conidial germi- nation and CAT fusion, whereas the solvent control (2% methanol) had no
effect (Fig. 1B and C). For example, at a concentration of7.5 pg DAPG ml -1, 68% of conidia germinated and 42% of these were involved
in one or more CAT fusion events, compared to 95% and 56%, respectively for the untreated control. From these data it was possible to
calculate an IC50 of ~9 pg DAPG ml -1 for both conidial germination and CAT fusion, which is comparable to the concentrations of DAPG
that inhibit other fungi (Kwak et al., 2009; Schouten et al., 2008, 2004). Given that some strains of P. fluorescens can generate
concentrations as high as 50100 pg DAPG ml-1 in culture (Delany et al., 2000), it is also a bio- logically relevant concentration.

3.2. DAPG treatment disrupts mitochondrial morphology

It has recently been shown that mitochondria are a target of DAPG toxicity in S. cerevisiae (Gleeson et al., 2010) and therefore the
effects of DAPG on mitochondria in N. crassa were investigated. To visualise changes in mitochondrial morphology, the fluorescent
styryl dye DASPMI, which preferentially stains mitochondria (Hickey et al., 2005), was added to N. crassa germlings, which were then
visualised using epifluorescence microscopy. In germlings
AB

.
G
e
r
<f m
in
100
40 at
o io
30
90 n
a) C
20
80 germination and CAT fusion by DAPG in the wild type. (A) Colony extension on agar plates
Fig. 1. Inhibition of colony extension rates, conidial A is normal on Vogels (-DAPG) and
T
70
reduced in the presence ofDAPG. No effect on extension
10 rate was observed in the presence of the solvent control (2% methanol). (B) Conidia
fu germinate and undergo cell fusion in
60
Vogels (-DAPG) but fail to do so when these processes are inhibited when 15 pg ml -1 DAPG is added (+DAPG). Scale bar: 5 pm. (C) Both germination
si and CAT fusion are inhibited
by DAPG in a dose-dependent manner. 2% CO 050 was 0used
O) Methanol in the7.5
MetOH treatment/DAPG
solvent control. [pg/ml]
9 11.4 13.2 15 on
6 D.M. Troppens et al./Fungal Genetics and Biology 56 (2013) 135-146

treated with this dye, mitochondria were either visible as tubular or

J
near spherical structures, with the tubular mitochondrial net-

X'/ _
works more evident at growing hyphal tips (Fig. 2A). When treated
with DAPG, a rapid and dramatic change was seen, with the fluo-
rescence signal becoming dispersed within minutes and some
apparent staining of membranes around large, spherical
structures (Fig. 2A). To ensure that the effects on mitochondrial

M
%
structure and the formation of these spherical structures were
not specific to the DASPMI dye, and to establish whether these
spherical bodies were nuclei, a second dye that stains

y
mitochondria, rhodamine B (Reu- ngpatthanaphong et al., 2003), + DAPG (rho B)
was used in a strain expressing his- tone H1-GFP to label its
nuclei (Freitag et al., 2004). Interestingly, mitochondria stained
with rhodamine B were all near spherical rather than tubular '"'tV -V
indicating that the dye exhibits cytotoxic ef- fects on
mitochondrial morphology. It was seen again, however, that DAPG + DAPG (DASPMI)
treatment led to a profound change in mitochondrial appearance,
with most fluorescence becoming associated with bright dot-like
structures, possibly representing fragmented mitochondria (Fig.
2B). The large spherical structures visible following DASPMI
+ DAPG (GFP)
treatment were also seen after rhodamine B treatment, indicating
that they are not an artefact associated with one specific dye.
Labelling with H1-GFP demonstrated that the large spherical
bodies were nuclei (compare Fig. 2A with Fig. 2B, lower panel).
The visibility of nuclei following treatment with DAPG was a result
- t

%
of some staining of nuclear membranes by DASPMI and
rhodamine B (Fig. 2A and B). When H1-GFP labelled cells were
stained with

- DAPG (DASPMI) - DAPG (rho B)


D.M. Troppens et al./Fungal Genetics and Biology 56 (2013) 135-146 7

Fig. 2. Effects of DAPG on mitochondrial morphology in conidial germlings of the causes a large increase in [Ca2+]c immediately following injection,
wild type imaged by widefield fluorescence microscopy. (A) Germlings with but [Ca2+]c quickly re- turns to its resting level, This is manifest
spherical and tubular mitochondria stained with DASPMI in Vogel's (-DAPG) and
as a highly reproducible Ca2+ signature and has been previously
9 min after the addition of DAPG (+DAPG). (B) Germlings of H1-GFP stained with
rhodamine B in Vogel's show mainly spherical mitochondria (-DAPG). In the described (Bencina et al., 2005; Binder et al., 2010; Nelson et al.,
presence of DAPG (+DAPG) (2 h exposure) spherical structures are visible, which 2004). When DAPG was injected, an additional and longer
match the GFP-labelled nuclei (arrows). The arrow in the lower pane of panel A transient increase in [Ca2+]c was observed following the first
indicates an example of the large, spherical structures seen after DAPG [Ca2+]c transient in response to mechanical perturbation (Fig. 4A).
treatment. rho B = rhodamine B Scale bar: 5 im.
The amplitude of this secondary [Ca2+]c transient was dose-
dependent and peaked approximately 2 min after addition of
rhodamine B in the absence of DAPG, the appearance of nuclei was DAPG before slowly returning the original [Ca 2+]c resting level
similar to that after DAPG treatment indicating that there is no after ~10 min. This effect appears to be quite specific to DAPG
evidence for an effect of DAPG on nuclear morphology (Fig. S1). and a structurally similar molecule, 6-methylsa-

3.3. DAPG treatment causes the loss of mitochondrial


membrane potential

To assess whether the DAPG-induced alterations in mitochon-


drial structure seen with DASPMI and rhodamine B were associ-
ated with a loss in mitochondrial membrane potential, a time
course experiment was performed using confocal microscopy. Be-
cause the potentiometric dye rhodamine B appeared to have some
cytotoxic effects on mitochondria, an alternative potentiometric
dye, rhodamine 123 (Hickey et al., 2005), was used. This dye,
which was rapidly taken up by cells and fluoresces only in
normally charged mitochondrial membranes, is commonly used
to measure mitochondrial membrane potential. When
mitochondria in un- treated conidial germlings were visualised
with rhodamine 123, they appeared spherical or tubular,
consistent with them being functional mitochondria (Fig. 3A).
When treated with 60 ig DAPG ml-1 the intensity of mitochondrial
fluorescence decreased and disappeared completely within 10
min (Fig. 3B), whereas no effect on mitochondrial fluorescence
intensity was observed in the control which was only treated with
Vogels liquid medium. Similar effects were observed using lower
concentrations of 15 ig DAPG ml-1 (data not shown). To further
investigate these findings, two known inhibitors of mitochondrial
function, antimy- cin A and CCCP, were used to treat conidial
germlings in the same manner. Antimycin A binds specifically to
the Complex III of the respiratory chain and inhibits electron
transfer, whereas CCCP is an uncoupler of oxidative
phosphorylation and ATP synthesis. When antimycin A was
added, a loss of mitochondrial fluorescence intensity was also
observed but this was less pronounced and slower than DAPG
treatment (data not shown). In contrast, the effects of CCCP were
very similar to those of DAPG with a rapid decrease of
mitochondrial fluorescence (Fig. 3C). These data are consistent
with DAPG acting as a chemiosmotic uncoupler and dis- rupting
mitochondrial membrane potential in a manner analogous to
CCCP. It is also possible that the loss of mitochondrial membrane
potential is a consequence of the effects of DAPG on
mitochondrial structure.

3.4. DAPG affects intracellular Ca2+ homeostasis

Several previous studies have suggested that DAPG may affect


multiple cellular processes and given the central importance of
Ca2+ homeostasis in filamentous fungi (Binder et al., 2010;
Nelson et al., 2004), it was decided to determine whether DAPG
treatment had any effects on Ca 2+ levels in conidial germlings.
Cytosolic free Ca2+ ([Ca2+]c) concentrations can be routinely
measured in conidial germlings of N. crassa germlings using the
heterologously-ex- pressed, bioluminescent Ca2+-reporter
aequorin (Binder et al., 2010). In this assay, test compounds in
Vogels liquid medium are injected into the wells of a 96-well plate
containing conidial germlings. This mechanical perturbation
8 D.M. Troppens et al./Fungal Genetics and Biology 56 (2013) 135-146

before

1 min

10 min

V
*V
'K

Control DAPG
CCCP
Fig. 3. DAPG-induced loss of mitochondrial membrane potential in conidial germlings. Time course experiment using confocal microscopy to show brightfield images (top) and
fluorescence images (below) before, and 1 min and 10 min after treatment. (A) Treatment with liquid Vogel's medium has no effect on mitochondrial fluorescence intensity. (B)
Treatment with DAPG leads to rapid loss of mitochondrial fluorescence within 10 min similar to (C) treatment with CCCP, an uncoupler of ATP synthesis, scale bar: 5 pm.

licylate did not induce this secondary increase in [Ca 2+]c at equiv-
alent concentrations (Fig. 4B). To establish whether the increase
in [Ca2+]c was a consequence of Ca 2+ uptake from outside the cell
or Ca2+ release from intracellular stores, the Ca 2+ chelator BAPTA
was added prior to treatment with DAPG (Fig. 4C). In this case,
both the [Ca2+]c increase caused by mechanical perturbation and
that caused by DAPG were abolished, indicating that the DAPG-
induced [Ca2+]c transient requires an import of Ca2+ from the
external medium. To further explore this, the DAPG-induced
[Ca2+]c transient was assessed in deletion mutants of three genes
that encode the following Ca 2+ channel proteins: CCH-1, which is
a high affinity cytoplasmic membrane channel; FIG-1, which is a
low affinity cytoplasmic membrane channel, and YVC-1, which is
a vacuolar Ca2+ channel (Bencina et al., 2009; Zelter et al., 2004).
The mutants lacking CCH-1 or FIG-1 still exhibited the DAPG-
induced [Ca2+]c transient, although some alteration in the Ca 2+
signature was seen in both mutants (Fig. 4D and F). Deletion of
the yvc-1 gene had no discernible effect on the [Ca 2+]c transient
confirming that this channel plays no role in releasing Ca 2+ from
internal Ca2+ stores (Fig. 4E). Thus, although CCH-1 and FIG-1
may play some role in the import of Ca2+, neither of these
proteins is solely involved in the secondary [Ca2+]c transient seen
following DAPG treatment.
D.M. Troppens et al./Fungal Genetics and Biology 56 (2013) 135-146 9

To assess whether these Ca2+ channel proteins play a role in be a conse- quence of plasma membrane permeabilisation by
DAPG toxicity, the sensitivity of the Ca 2+ channel protein mutants DAPG, was determined. Conidia were treated with DAPG or
to DAPG was compared with the wild-type. In addition, a mutant ethanol in the presence of the live/dead cell dyes propidium
lacking another putative Ca2+ channel protein, namely MID-1, iodide (PI) and fluo- rescein diacetate (FDA) (Hickey et al., 2005).
was tested for its DAPG sensitivity. When mutants were grown on Whereas ethanol caused 100% of cells to be permeabilised, less
Vogels agar medium for 28 h in the presence of different con- than 3% of cells were permeabilised by DAPG treatment (Fig. S3),
centrations of DAPG, no significant increase in sensitivity or toler- indicating that the increase in [Ca2+]c is unlikely to be the result
ance to this compound was observed (Fig. 5). However, when of membrane perme- abilisation. The ~2 min lag between
themutants were grown for longer (48 h) in the presence of the injection of DAPG (when the secondary transient [Ca 2+]c increase
high- est concentration (15 pgmr1) of DAPG used in this was observed) and the subse- quent recovery of the [Ca 2+]c to its
experiment, increased sensitivity of all of the mutants was noted resting level is further evidence that the plasma membrane is not
with Dmid- 1 being the most sensitive (Fig. 6). Nevertheless, a link compromised by DAPG treatment.
between the DAPG-induced [Ca2+]c transient and DAPG toxicity
was not found: when cells were pre-treated with BAPTA to abolish
the [Ca2+]c transient (Fig. 4C), this did not prevent the cells from 4. Discussion
losing their rhodamine 123 fluorescence after DAPG was added
(Fig. S2). N. crassa displays a comparable level of sensitivity to DAPG as
various phytopathogenic fungi and budding yeast (Gleeson et al.,
2010; Kwak et al., 2009; Schouten et al., 2008, 2004), confirming
Given the effects that DAPG has on mitochondrial
that it is an appropriate model to assess DAPG mode-of-action
membranes, whether the DAPG-induced [Ca2+]c transient could
.

4. The effects of DAPG on [Ca2+]c in conidial germlings. (A) A secondary [Ca2+]c transient induced by DAPG is dose-dependent in the wild type. (B) The secondary [Ca 2+]c transient is specific to DAPG because it
is not induced by the structurally similar molecule, 6-methylsalicylate. (C) Treatment with the Ca 2+ chelator, BAPTA, demonstrates that the DAPG-induced increase in [Ca 2+]c originates from the external
medium. Lacle of Ca2+ channels. (D and F) The DAPG-induced, transient increases are dose-dependent in the cch-1, yvc-1 and fig-1 mutants, respectively, but their Ca2+ signatures differ slightly from those in
the wild type (compare with Fig. 4A).

e capacity, in a genetically tractable fungus with a large suite of cell biology and determinethe specific effects of DAPG on conidial germlings of N. crassa. We show
genetic tools, to measure impacts on fungal cell growth, development and that DAPG inhibits fungal growth by impairing mito- chondrial function. DAPG
physiology makes N. crassa an excellent experimental system for this purpose. In was also found to transiently elevate [Ca2+]c but no evidence linking this
this study, various phys- iological probes and mutant strains were employed to increased [Ca2+]c with DAPG toxicity was obtained.
response to DAPG indicating that this is likely to be a primary sub- cellular
target of the metabolite. The reduction in fluorescence de- tected with rhodamine
123 is consistent with a loss of mitochondrial membrane potential, similar to
that seen with agents that uncouple oxidative phosphorylation from ATP genera-
tion by dissipating the proton gradient (e.g. CCCP and 2,4 dinitro- phenol).
Indeed, it is notable, that DAPG shares structural and chemical properties, for
130 example, a phenolic moiety and lipophilic nature,
wt with known uncouplers (Draber et al., 1972;
120
A fig-1 Heytler, 1979; Tollenaere, 1973). Previously, a loss
110 A yvc-1 of mitochondrial membrane potential was
100 A cch-1
observed with the carbocyanine dye DiOCl6(3) in
the yeast S. cerevisiae (Gleeson et al., 2010) and
90 A mid-1
effects on other fungi and oomycetes would also
80 be consistent with the uncoupling of oxidative
70 phosphorylation and ATP synthesis (de Souza et
al., 2003; Schouten et al., 2008, 2004). Recent
60
biochemical studies in yeast found that the effects
50 of CCCP and DAPG on respiratory function were
40 similar, providing further support for the
30 suggestion that DAPG has proton ionophoric
activity (Troppens et al., 2013). The capacity to act
20 as an uncoupling agent also offers an explanation
10 0 for why DAPG inhibits the growth of diverse

5 10 15
DAPG [pg/mi]

Fig. 5. No inhibition of colony extensin in the wild type and Ca 2+ channel protein mutant strains over 28 h. The lack of the Ca 2+ channel proteins FIG-1, YVC-1, CCH-1 and MID-
1 in each the deletion mutants does not change the sensitivity to DAPG when grown in the presence of different concentrations of DAPG and incubated for 28 h.

wt control 15 |jg/ml DAPG


organisms that indude bacteria, fungi, oomycetes and helminths, as the
requirement to maintain electrochemical gradients to generate energy is uni-
versal. The precise means by which uncoupling occurs has not been established,
4.1. D though the structural similarity to dinitrophenol may suggest a mechanism of H+
APG inhibits
growth by transport across the inner mitochon- drial membrane. In addition to affecting
membrane potential, the structure and possibly integrity of the mitochondria
were altered by DAPG. Several studies in N. crassa have found a link between
mitochondrial morphology, cytoskeletal elements and growth (Prokisch et al.,
2000). For example ropy mutants which are defective in the dynein-dynactin
A fig-1
motor protein complex, exhibited altered microtubule organisation and have dot-
impairing like mitochondria in N. crassa (Minke et al., 1999), not dissimilar to those
observed in this study. In both S. cerevisiae and Schizosaccharomyces pombe the
requirement of specific outer membrane proteins for normal mitochondrial
morphology has been demonstrated (Burgess et al., 1994; Fritz et al., 2001). A
significant mitochondrial membrane potential is required for protein transfer into
A yvc-1 mitochondria (Schleyer et al., 1982) and it is possible that disrupting this mem-
mitochondrial brane potential triggers the morphological changes that are ob- served following
function DAPG treatment

Using three different


mitochondrial
Acch-1
probes,
detrimental
effects of DAPG
on mitochondrial
function were
A mid-1 apparent, with
the rapid
. In this study, it was found that, in response to DAPG, there is a transient
increase in the average [Ca2+]c in a population of conidial germlings of N. crassa.
This Ca2+ originates from outside the cell and as no significant membrane
Fig. 6. Inhibition of growth and signs of stress in colonies of the wild type and Ca 2+
channel protein mutant strains when grown for 48 h. Cultures were incubated for 48
h in the absence and in the presence of 15 pgmr 1 DAPG and visualised by
photography.
3contributes to a general stress-response that helps the cell cope with the
effects of DAPG. Although DAPG did not specifically affect the growth of the
4.2. 4.2. DAPG induces a transient [Ca2 3+]c increase Dcch-1, Dmid-1 or Dfig-1 mutants, it is notable that after 48 h of incubation in its
pres- ence, the mutant colonies as well as those of the wild-type were showing
signs of stress (Fig 6). In S. cerevisiae, Ca2+ is a well known second messenger
2The increase in [Ca +] 2
c may act as an intracellular signal that involved in mediating stress responses and it has been shown in fungal cells that
permeabilisation occurs, it must enter the cell via a channel or transporter of this still leaves the questions of how the DAPG activates the channel(s)
some kind. Although Ca2+ signalling and homeostasis are of fundamental responsible for Ca2+ uptake into germlings, what channel(s) is/are used to take
importance for many processes, Ca2+ uptake in fungi is little understood. cch- 1 up extracellular Ca2+, and what are the downstream consequences of the
Encodes a protein that is clearly recognisably a plasma mem- brane Ca 2+ increased [Ca2+]c?
channel, and many studies have demonstrated that Cch1p in the budding yeast
is a high affinity Ca2+ channel (HACS) that responds to stress, including ER In fungi, multiple distinct stimuli employ Ca2+ as a second mes- senger (Aiello
stress, and other stimuli (Courchesne et al., 2011; Fischer et al., 1997; Hong et et al., 2002; Araki et al., 2009; Binder et al., 2010; Greene et al., 2002; Kallies et
al., 2010; Liu et al., 2006). In addition, mid-1 encodes a protein that appears to al., 1998; Moser et al., 1996; Viladev- all et al., 2004) so understanding how
be sometimes associated with Cch1p (Hong et al., 2013) but that may also be particular responses are reg- ulated by changes in [Ca 2+] is a great challenge.
independently regulated - for example, it is some- times described as being part Some insights may come from plant and animal biology where the concept of
of a mechanosensitive channel (Lew et al., 2008; Peiter et al., 2005). A third gene, Ca2+ signatures has been developed further (Cheung et al., 2011; McAinsh and
fig-1, encodes a protein that is involved in Ca 2+ uptake during mating (Muller et Pittman, 2009; Webb et al., 1996). Depending on the spatial and temporal
al., 2003; Yang et al., 2011). Neither Mid1p nor Fig1p display any of the typical dynamics of the intracellular Ca 2+ signal, different phenotypic outcomes are
features associated with Ca2+ channels and it is not known whether they are possible. It seems likely that Ca 2+ signatures encoding information that can be
themselves transporters or regulators of true channel proteins. In S. cerevisiae, decoded by the signal-response machinery will also feature in fungi. In this
there are additional com- ponents of this Cch1p high affinity channel (Martin et regard, Ca2+ pulses with species and age specific signatures were recently
al., 2011), and one recent study suggests that at least one additional Ca 2+- observed at the subcellular level in filamentous fungi using the genetically
channel exists in that yeast (Groppi et al., 2011). Homologues of budding yeast encoded fluorescent Ca2+ reporter cameleon (Kim et al., 2012). However, the
CCH1, MID1 and FIG1 are found in filamentous fungi (Bencina et al., 2009; Zelter information (if any) encoded in these signatures was not determined. In relation
et al., 2004) and mutants defective in these proteins were used in this study to to the Ca2+ signature observed by aequorin luminescence measurements of cell
try to ascertain the signif- icance of changes in [Ca 2+]c levels for DAPG toxicity in popula- tions in the present study, the kinetics of the transient [Ca 2+]c increase
N. crassa, and specifically, whether the increase in [Ca2+]c levels exacerbated the that is induced by DAPG in N. crassa is interesting as the slow return its resting
effects of DAPG. The Dcch-1, Dmid-1 or Dfig-1 mutants all dis- played increased level is not typical of what has previously been observed in this fungus (Binder et
sensitivity to DAPG in terms of growth inhibition suggesting that, at least al., 2010).
individually, they are not involved in DAPG toxicity. Furthermore, none of the
Ca2+ channel mutants showed a markedly different Ca 2+ signature in response to 4.3. Implications for understanding the ecological role ofDAPG
DAPG treatment. In addition, preventing Ca 2+ uptake by chelating extracellular
Ca2+ (BAPTA addition) did not have any impact on the kinetics of mitochondrial This study identified two, possibly unlinked, effects of DAPG on fungi:
impairment. It was interesting that the increase in [Ca 2+]c still occurred in the antimicrobial activity by inhibition of mitochondrial func- tion and stimulation of
Dcch-1 and Dfig-1 mutants, indicating that a different system involved in Ca2+ intracellular Ca2+ uptake, thus affecting Ca2+ homeostasis. The mitochondrial
uptake from the external medium is induced by DAPG. Difficulties in generating impairment conforms to the traditional view of DAPG as an antibiotic produced
a Dmid-1 strain expressing the aequorin reporter (Chu & Read, unpublished by P. fluorescens as a defence against protozoan predation or to antagonise other
data) meant that this mutant could not be tested in the present study but given microbes and thereby gain a competitive advantage. In this re- spect, DAPG
that budding yeast Mid1p commonly works with Cch1p, the most likely would be broadly similar to a number of other micro- bial metabolites that impair
explanation would seem to be that an as- yet undescribed channel or transporter mitochondrial function. For example, the F(1)F(0) ATP synthase (inhibited by
mediates the DAPG-in- duced Ca2+ import across the plasma membrane. In oligomycin), complex III (inhibited by antimycin A, ilicicolin H, funiculosin and
summary, our results provide no clear link between DAPG-induced [Ca 2+]c in- strobilurin) and complex II (inhibited by atpenins) are all specifically impaired by
crease and DAPG toxicity although the growth of the Ca 2+ channel mutants was bacterial and fungal metabolites (Bartlett et al., 2002; Gutierrez- Cirlos et al.,
more sensitive to DAPG than the wild type. 2004; Miyadera et al., 2003; Slater, 1973; vonJagow and Link, 1986). It is far
from certain, however, that DAPG is pro- duced at sufficient levels in the
The DAPG-induced [Ca2+]c transient is not involved in DAPG toxicity and rhizosphere to mediate significant antibiotic effects, and, in fact, the pattern of
understanding its significance will require further work. There are three possible DAPG production is more reminiscent of a signal than an antibiotic, as DAPG
roles that this transient [Ca2+]c increase may serve: levels rise sharply in late exponential phase and then rapidly decline as the
molecule is turned over (Bottiglieri and Keel, 2006; Delany et al., 2000). The
endoplasmic reticulum (ER) stress activates the Cch1p- Mid1p channel possibility that DAPG serves as a signal is intriguing and in that regard, the
complex that acts to restore ER Ca 2+ levels (Hong et al., 2010, 2013). The finding that DAPG affects Ca 2+ homeostasis in a eukaryotic (fungal) cell is very
best-known mode-of-action of Ca2+ involves its binding to calmodulin, interesting because of the known importance of the Ca2+ second messenger in
which then inter- acts with the protein phosphatase calcineurin leading to plant-microbe interactions. Both the rhizobium-legume symbiosis and the sym-
activation of various target proteins, including transcription factors (Cyert, biosis between mycorrhizal fungi and diverse plants involves microbial signal-
2003; Kraus and Heitman, 2003). induced Ca2+oscillations in the plant cell (Downie, 2010; Navazio et al., 2007).
The effect of the Ca2+ oscillations is to activate patterns of gene expression that
(2) The [Ca2+]c transient may be a consequence of the effects of DAPG on are required for develop- ment of the symbiotic state. It has also been reported
mitochondria. Although there are no data from fungi, one study with sea that a [Ca2+]c transient in plant cells is part of the plant-pathogen interaction
urchin sperm found that impairing mitochondrial function led to an that is required for the activation of systemic resistance by plants (Lecourieux et
uptake of extracellular Ca2+ in a mechanism that commenced with al., 2006). Although speculative at this stage, a role for DAPG in Ca 2+ signalling
mitochondrial Ca2+ release (Ardon et al., 2009). This is a possibility that is an intriguing concept that should be further explored.
should be explored but the differences between the Ca 2+ signal-
ling/homeostatic machinery of animal and fungal cells (Bor- kovich et al., 4.4. Conclusions
2004; Galagan et al., 2003), and the rather unique cell-type used in that
study (the sea urchin sperm cell) advise caution in extrapolating between By assessing the effects of DAPG on N. crassa, we have now established that
animal and fungal systems. the antibiotic activity of DAPG is a consequence of its capacity to act as an
uncoupler in the mitochondrion. This has important implications for the
(3) The transient [Ca2+]c increase could be a distinct response induced by development of biological control inoculants as it confirms that the activity of
DAPG via an unknown mechanism. This option is supported by the DAPG-producing strains will be broad spectrum. Furthermore, this is a distinct
observation that the Cch1p channel is not required for Ca 2+ uptake in the mode-of-action from other antifungal agents. It also suggests that resistance will
response of N. crassa to DAPG (Fig. 4). be a rare event as the mitochondrial membrane is a not easily altered target. It
will be interesting to see if some of the natural Fusarium isolates that show
Whichever of these three possible roles that the DAPG-induced [Ca 2+]c has, reduced susceptibility to DAPG (Schouten et al., 2008) are also more tolerant to
other uncou- pling agents. It is also noteworthy that anti-cancer and pro-apop- Delany, I. et al., 2000. Regulation of production of the antifungal metabolite 2,4-
diacetylphloroglucinol in Pseudomonas fluorescens F113: genetic analysis of phlF as a
totic effects have been attributed to various uncoupling agents (Han et al., 2008;
transcriptional repressor. Microbiology 146 (Pt 2), 537-543.
Pardo-Andreu et al., 2011). In this study, a second physiological effect of DAPG Demain, A.L., 2009. Antibiotics: natural products essential to human health. Med. Res. Rev. 29, 821-
on fungi, namely alteration of intracel- lular Ca 2+ homeostasis by inducing 842.
uptake of Ca2+ from the extra- cellular environment via an unknown Downie, J.A., 2010. The roles ofextracellularproteins, polysaccharides and signals in the interactions
of rhizobia with legume roots. FEMS Microbiol. Rev. 34, 150170.
channel/transport system, has also been observed. This raises the possibility Draber, W. et al., 1972. Quantitative structure-activity studies of hydrazones, uncouplers of oxidative
that DAPG acts as a bacterial signal that is perceived by eukaryotic cells. If so, phosphorylation. Z. Naturforsch B. 27,159-171.
this could have important ramifications for the interaction of DAPG- producing Duan, K. et al., 2009. Chemical interactions between organisms in microbial communities. Contrib.
Microbiol. 16,1-17.
bacteria with fungi, protozoa and/or plants. Genetically tractable fungi like N.
Ermolayeva, E., Sanders, D., 1995. Mechanism of pyrithione-induced membrane depolarization in
crassa, or indeed S. cerevisiae, offer a means to further explore all of these Neurospora crassa. Appl. Environ. Microbiol. 61, 3385-3390.
possibilities. Fenton, A.M. et al., 1992. Exploitation of gene(s) involved in 2,4- diacetylphloroglucinol biosynthesis
to confer a new biocontrol capability to a Pseudomonas strain. Appl. Environ. Microbiol. 58,
3873-3878.
nowledgments Fischer, M. et al., 1997. The Saccharomyces cerevisiae CCH1 gene is involved in calcium influx and
mating. FEBS Lett. 419, 259-262.
Freitag, M. et al., 2004. GFP as a tool to analyze the organization, dynamics and function of nuclei
Funding support for this Project was provided by a Biotechnol- ogy and
and microtubules in Neurospora crassa. Fungal Genet. Biol. 41, 897-910.
Biological Sciences Research (BBSRC) Council Grant (15/ P18594) to N.D.R. and Fritz, S. et al., 2001. Connection of the mitochondrial outer and inner membranes by Fzo1 is critical
by Grants from the Science Foundation Ire- land (08/RFP/GEN1295), European for organellar fusion. J. Cell Biol. 152, 683-692.
Commission (O36314) and the Irish Department of Agriculture and Marine Galagan, J.E. et al., 2003. The genome sequence of the filamentous fungus Neurospora crassa. Nature
422, 859-868.
(BEAU/BIOD/01) to J.P.M. and F.OG. Technical support from Pat Higgins is Gleeson, O. et al., 2010. The Pseudomonas fluorescens secondary metabolite 2,4
gratefully acknowledged. diacetylphloroglucinol impairs mitochondrial function in Saccharomyces cerevisiae. Anton. Van
Leeuw. 97, 261-273.
Goh, E.B. et al., 2002. Transcriptional modulation of bacterial gene expression by subinhibitory
pendix A. Supplementary material concentrations of antibiotics. Proc. Natl. Acad. Sci. USA 99, 17025-17030.
Greene, V. et al., 2002. Oxidative stress-induced calcium signalling in Aspergillus nidulans. Cell.
Signal. 14, 437-443.
Supplementary data associated with this article can be found, in the online
Groppi, S. et al., 2011. Glucose-induced calcium influx in budding yeast involves a novel calcium
version, at http://dx.doi.org/10.1016Zj.fgb.2013.04.006. transport system and can activate calcineurin. Cell Calcium 49, 376-386.
References Gross, H., Loper, J.E., 2009. Genomics of secondary metabolite production by Pseudomonas spp. Nat.
Prod. Rep. 26, 1408-1446.
Achkar, J. et al., 2005. Biosynthesis of phloroglucinol. J. Am. Chem. Soc. 127, 53325333.
Gutierrez-Cirlos, E.B. et al., 2004. Inhibition of the yeast cytochrome bc1 complex by ilicicolin H, a
Aiello, D.P. et al., 2002. Intracellular glucose 1-phosphate and glucose 6-phosphate levels modulate
novel inhibitor that acts at the Qn site of the bc1 complex. J. Biol. Chem. 279, 8708-8714.
Ca2+ homeostasis in Saccharomyces cerevisiae. J. Biol. Chem. 277, 45751-45758.
Haas, D., Defago, G., 2005. Biological control of soil-borne pathogens by fluorescent pseudomonads.
Araki, Y. et al., 2009. Ethanol stress stimulates the Ca 2+-mediated calcineurin/Crz1 pathway in
Nat. Rev. Microbiol. 3, 307-319.
Saccharomyces cerevisiae. J. Biosci. Bioeng. 107,1-6.
Han, Y.H. et al., 2008. 2,4-Dinitrophenol induces G1 phase arrest and apoptosis in human
Ardon, F. et al., 2009. Mitochondrial inhibitors activate influx of external Ca(2+) in sea urchin sperm.
pulmonary adenocarcinoma Calu-6 cells. Toxicol. In Vitro 22, 659-670.
Biochim. Biophys. Acta 1787, 15-24.
Heytler, P.G., 1979. Uncouplers of oxidative phosphorylation. Methods Enzymol. 55, 462-42.
Bangera, M.G., Thomashow, L.S., 1996. Characterization of a genomic locus required for synthesis of
Hickey, P.C. et al., 2005. Live-cell imaging of filamentous fungi using vital fluorescent dyes. In:
the antibiotic 2,4-diacetylphloroglucinol by the biological control agent Pseudomonas fluorescens
Methods in Microbiology. Microbial Imaging, vol. 34. Elsevier, Amsterdam, pp. 63-87.
Q2-87. Mol. Plant Microbe Interact. 9, 83-90.
Hong, M.P. et al., 2010. Cch1 restores intracellular Ca2+ in fungal cells during endoplasmic
Bartlett, D.W. et al., 2002. The strobilurin fungicides. Pest Manage. Sci. 58,649-662.
reticulum stress. J. Biol. Chem. 285, 10951-10958.
Bencina, M. et al., 2005. Cross-talk between cAMP and calcium signalling in Aspergillus niger. Mol.
Hong, M.P. et al., 2013. Activity of the calcium channel pore Cch1 is dependent on a modulatory
Microbiol. 56, 268-281.
region of the subunit Mid1 in Cryptococcus neoformans. Eukaryot. Cell. 12, 142-150.
Bencina, M. et al., 2009. A comparative genomic analysis of calcium and proton
Iavicoli, A. et al., 2003. Induced systemic resistance in Arabidopsis thaliana in response to root
signaling/homeostasis in Aspergillus species. Fungal Genet. Biol. 46 (Suppl. 1), S93-S104.
inoculation with Pseudomonas fluorescens CHA0. Mol. Plant Microbe Interact. 16, 851-858.
Berepiki, A. et al., 2010. F-actin dynamics in Neurospora crassa. Eukaryot. Cell 9, 547-557.
Jones Jr., S.K., Bennett, R.J., 2011. Fungal mating pheromones: choreographing the dating game.
Binder, U. et al., 2010. The antifungal activity of the Penicillium chrysogenum protein PAF disrupts
Fungal Genet. Biol. 48, 668-676.
calcium homeostasis in Neurospora crassa. Eukaryot. Cell 9,13741382.
Jousset, A. et al., 2006. Secondary metabolites help biocontrol strain Pseudomonas fluorescens CHA0
Bloemberg, G.V., Lugtenberg, B.J., 2001. Molecular basis of plant growth promotion and biocontrol
to escape protozoan grazing. Appl. Environ. Microbiol. 72, 7083-7090.
by rhizobacteria. Curr. Opin. Plant Biol. 4, 343-350.
Kallies, A. et al., 1998. Heat shock effects on second messenger systems of Neurospora crassa. Arch.
Borkovich, K.A. et al., 2004. Lessons from the genome sequence of Neurospora crassa: tracing the
Microbiol. 170,191-200.
path from genomic blueprint to multicellular organism. Microbiol. Mol. Biol. Rev. 68, 1-108.
Keel, C. et al., 1992. Suppression of root diseases by Pseudomonas fluorescens CHA0: Importance of
Bottiglieri, M., Keel, C., 2006. Characterization of PhlG, a hydrolase that specifically degrades the
the secondary metabolite 2,4-diacetylphloroglucinol. Mol. Plant Microbe Interact. 5, 4-13.
antifungal compound 2,4-diacetylphloroglucinol in the biocontrol agent Pseudomonas
Keel, C. et al., 1996. Conservation of the 2,4-diacetylphloroglucinol biosynthesis locus among
fluorescens CHA0. Appl. Environ. Microbiol. 72, 418-427.
fluorescent Pseudomonas strains from diverse geographic locations. Appl. Environ. Microbiol. 62,
Burgess, S.M. et al., 1994. MMM1 encodes a mitochondrial outer membrane protein essential for
552-563.
establishing and maintaining the structure of yeast mitochondria. J. Cell Biol. 126, 1375-1391.
Kim, H.S. et al., 2012. Expression of the Cameleon calcium biosensor in fungi reveals distinct Ca(2+)
Calabrese, E.J. et al., 2008. Hormesis predicts low-dose responses better than threshold models. Int.
signatures associated with polarized growth, development, and pathogenesis. Fungal Genet. Biol.
J. Toxicol. 27, 369-378.
49, 589-601.
Challis, G.L., Hopwood, D.A., 2003. Synergy and contingency as driving forces for the evolution of
Kraus, P.R., Heitman, J., 2003. Coping with stress: calmodulin and calcineurin in model and
multiple secondary metabolite production by Streptomyces species. Proc. Natl. Acad. Sci. USA
pathogenic fungi. Biochem. Biophys. Res. Commun. 311, 1151-1157.
100 (Suppl 2), 14555-14561.
Kuroda, H. et al., 2007. Subinhibitory concentrations of beta-lactam induce haemolytic activity in
Cheung, C.Y. et al., 2011. Visualization, characterization and modulation of calcium signaling during
Staphylococcus aureus through the SaeRS two- component system. FEMS Microbiol. Lett. 268,
the development of slow muscle cells in intact zebrafish embryos. Int. J. Dev. Biol. 55, 153-174.
98-105.
Cornelis, P., 2010. Iron uptake and metabolism in pseudomonads. Appl. Microbiol. Biotechnol. 86,
Kwak, Y.S. et al., 2009. Diversity, virulence, and 2,4-diacetylphloroglucinol sensitivity of
1637-1645.
Gaeumannomyces graminis var. tritici isolates from Washington state. Phytopathology 99, 472-
Courchesne, W.E. et al., 2011. Ethanol induces calcium influx via the Cch1-Mid1 transporter in
479.
Saccharomyces cerevisiae. Arch. Microbiol. 193, 323-334.
Kwak, Y.S. et al., 2010. Saccharomyces cerevisiae genome-wide mutant screen for sensitivity to 2,4-
Cronin, D. et al., 1997a. Role of 2,4-diacetylphloroglucinol in the interactions of the biocontrol
diacetylphloroglucinol, an antibiotic produced by Pseudomonas fluorescens. Appl. Environ.
pseudomonad strain F113 with the potato cyst nematode Globodera rostochiensis. Appl. Environ.
Microbiol. 77, 1770-1776.
Microbiol. 63,1357-1361.
Laville, J. et al., 1992. Global control in Pseudomonas fluorescens mediating antibiotic synthesis and
Cronin, D. et al., 1997b. Ecological interaction of a biocontrol Pseudomonas fluorescens strain
suppression of black root rot of tobacco. Proc. Natl. Acad. Sci. USA 89, 1562-1566.
producing 2,4-diacetylphloroglucinol with the soft rot potato pathogen Erwinia carotovora subsp.
Lecourieux, D. et al., 2006. Calcium in plant defence-signalling pathways. New Phytol. 171, 249-269.
atroseptica. FEMS Microbiol. Ecol. 23, 95-106.
Lew, R.R. et al., 2008. Phenotype of a mechanosensitive channel mutant, mid-1, in a filamentous
Cummins, J. et al., 2009. Subinhibitory concentrations of the cationic antimicrobial peptide colistin
fungus, Neurospora crassa. Eukaryot. Cell 7, 647-655.
induce the pseudomonas quinolone signal in Pseudomonas aeruginosa. Microbiology 155, 2826-
Liu, M. et al., 2006. Cch1 mediates calcium entry in Cryptococcus neoformans and is essential in low-
2837.
calcium environments. Eukaryot. Cell 5, 1788-1796.
Cyert, M.S., 2003. Calcineurin signaling in Saccharomyces cerevisiae: how yeast go crazy in response
Marris, P. 2007. Ca2+-signalling in response to mechanical perturbation and hypo- osmotic shock in
to stress. Biochem. Biophys. Res. Commun. 311, 1143-1150.
Neurospora crassa. Ph.D. thesis. University of Edinburgh, Edinburgh, UK.
Davis, R.H., 2000. Neurospora: Contributions of a Model Organism. Oxford University Press.
Martin, D.C. et al., 2011. New regulators of a high affinity Ca 2+ influx system revealed through a
Davis, R.H., Perkins, D.D., 2002. Timeline: Neurospora: a model of model microbes. Nat. Rev. Genet.
genome-wide screen in yeast. J. Biol. Chem. 286, 1074410754.
3, 397-403.
Mazzola, M. et al., 1995. Variation in sensitivity of Gaeumannomyces graminis to antibiotics produced
de Souza, J.T. et al., 2003. Effect of 2,4-diacetylphloroglucinol on Pythium: cellular responses and
by fluorescent Pseudomonas spp. and effect on biological control of take-all of wheat. Appl.
variation in sensitivity among propagules and species. Phytopathology 93, 966-975.
Environ. Microbiol. 61, 2554-2559. broad-spectrum antibiotic produced by Pseudomonas fluorescens. Mol. Plant Microbe Interact.
McAinsh, M.R., Pittman, J.K., 2009. Shaping the calcium signature. New Phytol. 181, 275-294. 17,1201-1211.
Mesquita, A. et al., 2010. Caloric restriction or catalase inactivation extends yeast chronological Schouten, A. et al., 2008. Involvement of the ABC transporter BcAtrB and the laccase BcLCC2 in
lifespan by inducing H2O2 and superoxide dismutase activity. Proc. Natl. Acad. Sci. USA 107, defence of Botrytis cinerea against the broad-spectrum antibiotic 2,4- diacetylphloroglucinol.
15123-15128. Environ. Microbiol. 10,1145-1157.
Minke, P.F. et al., 1999. Microscopic analysis of Neurospora ropy mutants defective in nuclear Selker, E.U., 1998. Trichostatin A causes selective loss of DNA methylation in Neurospora. Proc. Natl.
distribution. Fungal Genet. Biol. 28, 55-67. Acad. Sci. USA 95, 9430-9435.
Miyadera, H. et al., 2003. Atpenins, potent and specific inhibitors of mitochondrial complex II Shanahan, P. et al., 1993. Liquid chromatographic assay of microbially derived phloroglucinol
(succinate-ubiquinone oxidoreductase). Proc. Natl. Acad. Sci. USA 100, 473-477. antibiotics for establishing the biosynthetic route to production, and the factors affecting their
Morrissey, J.P. et al., 2002. Exploitation of genetically modified inoculants for industrial ecology regulation. Anal. Chim. Acta. 272, 271-277.
applications. Anton. Van Leeuw. 81, 599-606. Shen, L. et al., 2008. Modulation of secreted virulence factor genes by subinhibitory concentrations of
Morrissey, J.P. et al., 2004. Are microbes at the root of a solution to world food production? Rational antibiotics in Pseudomonas aeruginosa. J. Microbiol. 46, 441447.
exploitation of interactions between microbes and plants can help to transform agriculture. Siddiqui, I.A., Shaukat, S.S., 2004. Trichoderma harzianum enhances the production of nematicidal
EMBO Rep. 5, 922-926. compounds in vitro and improves biocontrol of Meloidogyne javanica by Pseudomonas fluorescens
Moser, M.J. et al., 1996. Ca2+-calmodulin promotes survival of pheromone-induced growth arrest by in tomato. Lett. Appl. Microbiol. 38, 169175.
activation of calcineurin and Ca2+-calmodulin-dependent protein kinase. Mol. Cell Biol. 16, 4824- Slater, E.C., 1973. The mechanism of action of the respiratory inhibitor, antimycin. Biochim. Biophys.
4831. Acta. 301, 129-154.
Muller, E.M. et al., 2003. Fig 1p facilitates Ca2+ influx and cell fusion during mating of Saccharomyces Spelbrink, R.G. et al., 2004. Differential antifungal and calcium channel-blocking activity among
cerevisiae. J. Biol. Chem. 278, 38461-38469. structurally related plant defensins. Plant Physiol. 135, 20552067.
Munoz, A. et al., 2012. Concentration-dependent mechanisms of cell penetration and killing by the de Stein, T., 2005. Bacillus subtilis antibiotics: structures, syntheses and specific functions. Mol.
novo designed antifungal hexapeptide PAF26. Mol. Microbiol.. Microbiol. 56, 845-857.
Munoz, A. et al., 2013. Two functional motifs define the interaction, internalization and toxicity of the Thevissen, K. et al., 1999. Permeabilization of fungal membranes by plant defensins inhibits fungal
cell-penetrating antifungal peptide PAF26 on fungal cells. PLoS One 8, e54813. growth. Appl. Environ. Microbiol. 65, 5451-5458.
Navazio, L. et al., 2007. A diffusible signal from arbuscular mycorrhizal fungi elicits a transient Tollenaere, J.P., 1973. Structure-activity relationships of three groups of uncouplers of oxidative
cytosolic calcium elevation in host plant cells. Plant Physiol. 144, 673-681. phosphorylation: salicylanilides, 2-trifluoromethylbenzimidazoles, and phenols. J. Med. Chem.
Nelson, G. et al., 2004. Calcium measurement in living filamentous fungi expressing codon-optimized 16, 791-796.
aequorin. Mol. Microbiol. 52, 1437-1450. Troppens, D.M. et al., 2013. Genome-wide investigation of cellular targets and mode of action of the
Ng, W.L., Bassler, B.L., 2009. Bacterial quorum-sensing network architectures. Annu. Rev. Genet. 43, antifungal bacterial metabolite 2,4-diacetylphloroglucinol in Saccharomyces cerevisiae. FEMS
197-222. Yeast Res.. http://dx.doi.org/10.1111/1567- 1364.12037.
Palma-Guerrero, J. et al., 2009. Chitosan permeabilizes the plasma membrane and kills cells of Verhagen, B.W. et al., 2010. Pseudomonas spp.-induced systemic resistance to Botrytis cinerea is
Neurospora crassa in an energy dependent manner. Fungal Genet. Biol. 46, 585-594. associated with induction and priming of defence responses in grapevine. J. Exp. Bot. 61, 249-
Palma-Guerrero, J. et al., 2010. Membrane fluidity determines sensitivity of filamentous fungi to 260.
chitosan. Mol. Microbiol. 75, 1021-1032. Viladevall, L. et al., 2004. Characterization of the calcium-mediated response to alkaline stress in
Pardo-Andreu, G.L. et al., 2011. The anti-cancer agent nemorosone is a new potent protonophoric Saccharomyces cerevisiae. J. Biol. Chem. 279, 4361443624.
mitochondrial uncoupler. Mitochondrion 11, 255-263. Vincent, M.N. et al., 1991. Genetic analysis of the antifungal activity of a soilborne Pseudomonas
Peiter, E. et al., 2005. The Saccharomyces cerevisiae Ca2+ channel Cch1pMid1p is essential for aureofaciens strain. Appl. Environ. Microbiol. 57, 2928-2934.
tolerance to cold stress and iron toxicity. FEBS Lett. 579, 56975703. von Jagow, G., Link, T.A., 1986. Use of specific inhibitors on the mitochondrial bc1 complex. Methods
Pereira, R.C., Said, S., 2009. Alterations in growth and branching of Neurospora crassa caused by Enzymol. 126, 253-271.
sub-inhibitory concentrations of antifungal agents. Rev. Argent. Microbiol. 41, 39-44. Webb, A.A.R. et al., 1996. Calcium ions as intracellular second messengers in higher plants. Adv. Bot.
Prokisch, H. et al., 2000. Role of MMM1 in maintaining mitochondrial morphology in Neurospora Res. 22, 45-96.
crassa. Mol. Biol. Cell. 11, 2961-2971. Weller, D.M. et al., 2007. Role of 2,4-diacetylphloroglucinol-producing fluorescent Pseudomonas spp.
Raaijmakers, J.M. et al., 2002. Antibiotic production by bacterial biocontrol agents. Anton. Van in the defense of plant roots. Plant Biol. (Stuttg) 9, 4-20.
Leeuw. 81, 537-547. Williams, P., Camara, M., 2009. Quorum sensing and environmental adaptation in Pseudomonas
Raaijmakers, J.M. et al., 2010. Natural functions of lipopeptides from Bacillus and Pseudomonas: aeruginosa: a tale of regulatory networks and multifunctional signal molecules. Curr. Opin.
more than surfactants and antibiotics. FEMS Microbiol. Rev. 34, 1037-1062. Microbiol. 12,182-191.
Reungpatthanaphong, P. et al., 2003. Rhodamine B as a mitochondrial probe for measurement and Yang, M. et al., 2011. Fig 1 facilitates calcium influx and localizes to membranes destined to undergo
monitoring of mitochondrial membrane potential in drug- sensitive and -resistant cells. J. fusion during mating in Candida albicans. Eukaryot. Cell 10, 435-444.
Biochem. Biophys. Methods 57, 1-16. Yim, G. et al., 2006. The truth about antibiotics. Int. J. Med. Microbiol. 296,163-170.
Roca, M.G. et al., 2005. Cell biology of conidial anastomosis tubes in Neurospora crassa. Eukaryot. Yim, G. et al., 2007. Antibiotics as signalling molecules. Philos. Trans. R. Soc. London, B Biol. Sci.
Cell 4, 911-919. 362, 1195-1200.
Schleyer, M. et al., 1982. Requirement of a membrane potential for the posttranslational transfer of Zelter, A. et al., 2004. A comparative genomic analysis of the calcium signaling machinery in
proteins into mitochondria. Eur. J. Biochem. 125, 109-116. Neurospora crassa, Magnaporthe grisea, and Saccharomyces cerevisiae. Fungal Genet. Biol. 41,
Schouten, A. et al., 2004. Defense responses of Fusarium oxysporum to 2,4- diacetylphloroglucinol, a 827-841.

Você também pode gostar