ArticlePDF Available

Evidence for production of the phytohormone indole-3-acetic acid by cyanobacteria

Authors:

Abstract and Figures

The ability of cyanobacteria to produce the phytohormone indole-3-acetic acid (IAA) was demonstrated. A colorimetric (Salkowski) screening of 34 free-living and symbiotically competent cyanobacteria, that represent all morphotypes from the unicellular to the highly differentiated, showed that auxin-like compounds were released by about 38% of the free-living as compared to 83% of the symbiotic isolates. The endogenous accumulation and release of IAA were confirmed immunologically (ELISA) using an anti-IAA antibody on 10 of the Salkowski-positive strains, and the chemical authenticity of IAA was further verified by chemical characterization using gas chromatography-mass spectrometry in Nostoc PCC 9229 (isolated from the angiosperm Gunnera) and in Nostoc 268 (free-living). Addition of the putative IAA precursor tryptophan enhanced IAA accumulation in cell extracts and supernatants. As the genome of the symbiotically competent Nostoc PCC 73102 contains homologues of key enzymes of the indole-3-pyruvic acid pathway, a transaminase and indolepyruvate decarboxylase (IpdC), the putative ipdC gene from this cyanobacterium was cloned and used in Southern blot analysis. Out of 11 cyanobacterial strains responding positively in the Salkowski/ELISA test, ipdC homologues were found in 4. A constitutive and possibly tryptophan-dependent production of IAA via the indole-3-pyruvic acid pathway is therefore suggested. The possible role of IAA in cyanobacteria in general and in their interactions with plants is discussed.
Content may be subject to copyright.
ORIGINAL ARTICLE
Elena Sergeeva Æ Anton Liaimer Æ Birgitta Bergman
Evidence for production of the phytohormone indole-3-acetic
acid by cyanobacteria
Received: 8 October 2001 / Accepted: 11 January 2002 / Published online: 22 March 2002
Springer-Verlag 2002
Abstract The ability of cyanobacteria to produce the
phytohormone indole-3-acetic acid (IAA) was demon-
strated. A colorimetric (Salkowski) screening of 34 free-
living and symbiotically competent cyanobacteria, that
represent all morphotypes from the unicellular to the
highly differentiated, showed that auxin-like compounds
were released by about 38% of the free-living as com-
pared to 83% of the symbiotic isolates. The endogenous
accumulation and release of IAA were confirmed im-
munologically (ELISA) using an anti-IAA antibody on
10 of the Salkowski-positive strains, and the chemical
authenticity of IAA was further verified by chemical
characterization using gas chromatography-mass spect-
rometry in Nostoc PCC 9229 (isolated from the angio-
sperm Gunnera) and in Nostoc 268 (free-living).
Addition of the putative IAA precursor tryptophan en-
hanced IAA accumulation in cell extracts and superna-
tants. As the genome of the symbiotically competent
Nostoc PCC 73102 contains homologues of key enzymes
of the indole-3-pyruvic acid pathway, a transaminase
and indolepyruvate decarboxylase (IpdC), the putative
ipdC gene from this cyanobacterium was cloned and
used in Southern blot analysis. Out of 11 cyanobacterial
strains responding positively in the Salkowski/ELISA
test, ipdC homologues were found in 4. A constitutive
and possibly tryptophan-dependent production of IAA
via the indole-3-pyruvic acid pathway is therefore sug-
gested. The possible role of IAA in cyanobacteria in
general and in their interactions with plants is discussed.
Keywords Cyanobacterium Æ Indole-3-acetic acid Æ
Indolepyruvate decarboxylase Æ Plant
symbiosis Æ Tryptophan
Abbreviations Chl a: chlorophyll a Æ ELISA: enzyme-
linked immunosorbent assay Æ IAA: indole-3-acetic
acid Æ IpdC: indolepyruvate decarboxylase Æ TPP:
thiamine pyrophosphate Æ Trp: tryptophan
Introduction
Cyanobacteria are prokaryotic photoautotrophs of
evolutionary significance (Schopf 2000). They rely on an
oxygenic photosynthesis involving chlorophyll, photo-
system I/II electron transport and CO
2
fixation (Fay
1992). Cyanobacteria range in their morphology from
unicellular to differentiated, and branched types and are
widespread in nature.
Cyanobacteria also constitute a direct evolutionary
link between pro- and eukaryotes as cyanobacteria were
progenitors of chloroplasts (Douglas 1994). The en-
gulfment of a cyanobacterium by a primitive pigment-
less ancestral eukaryote led to the genesis of plants with
photosynthetic chloroplasts. Some cyanobacteria, par-
ticularly within the genus Nostoc, still enter into intimate
relationships with some eukaryotic organisms repre-
senting all plant divisions (Rai et al. 2000). Hosts are
found within liver- and hornworts, a water-fern (Azolla),
cycads and one angiosperm, as well as some fungi
(primarily lichenized fungi) and protists. The primary
benefit for the plants may be ascribed to the nitrogen-
fixation capacity of the cyanobiont, as most of the fixed
nitrogen is transferred to the host. However, the in-
volvement of other cyanobacterial molecules promoting
growth and development of host plants is also a possi-
bility in cyanobacterial–plant interactions, although
none has been identified.
Phytohormones play an important role as signals and
regulators of growth and development in plants. Auxins,
among them in particular indole-3-acetic acid (IAA), are
the most studied plant growth regulators, and this in-
cludes physiological, biochemical and genetic aspects.
The ability to produce the phytohormones is often
considered as a trait of the plant kingdom. However,
Planta (2002) 215: 229–238
DOI 10.1007/s00425-002-0749-x
E. Sergeeva Æ A. Liaimer Æ B. Bergman (&)
Department of Botany,
Stockholm University, 10691 Stockholm, Sweden
E-mail: bergmanb@botan.su.se
Fax: +46-8-165525
production of phytohormones is also widespread among
soil and plant-associated prokaryotes (Costacurta and
Vanderleyden 1995). Phytohormones produced by
plant-associated bacteria are implicated as key deter-
minants in stimulation of plant growth, in plant patho-
genesis and in associative or plant–microbe symbiotic
interactions (Christiansen-Weniger 1998). As some
plant-interacting bacteria release high levels of IAA,
they may locally change the endogenous phytohormone
balance of the host and trigger responses, the nature of
which depends on the phytohormone concentration
generated, type of host and its sensitivity to hormones
(Glick et al. 1999). Known plant growth-promoting
bacteria include species within genera such as Herbas-
pirillum, Rhizobium and Azospirillum, interacting with
plant roots, while for instance Pseudomonas, Erwinia
and Agrobacterium elicit galls or tumours on plants.
Possibly, phytohormone release provides plant-inter-
acting bacteria with a tool by which to promote plant
cell division, growth and nutrient release and thereby
their own growth. Genetic mechanisms underlying IAA
biosynthesis and its regulation have been studied in
Pseudomonas, Agrobacterium, Rhizobium, Bradyrhizobi-
um and Azospirillum, and multiple pathways for IAA
synthesis have been identified, including both trypto-
phan-dependent and independent pathways (Glick et al.
1999).
Up to now, there has been no unequivocal identifi-
cation of phytohormone production in cyanobacteria.
The aim of the present study was to examine the po-
tential in cyanobacteria to produce the phytohormone
IAA. This was done by screening for endogenous IAA
levels and IAA release in a range of free-living cyano-
bacteria, belonging to various morphotypes, and also in
symbiotically competent cyanobacteria. We demonstrate
that cyanobacteria, and in particular symbiotic isolates,
are capable of both accumulating and releasing IAA,
and suggest that IAA accumulation is stimulated by
exogenous tryptophan and may proceed via the indole-
3-pyruvic acid pathway.
Materials and methods
Organisms and growth conditions
The 34 species of cyanobacteria examined were obtained from the
sources given in Table 1 and all are maintained in our culture
collection (Department of Botany at Stockholm University). The
cyanobacteria were re-inoculated every second week and were
grown axenically in either nitrogen-free BG-11
0
medium (for ni-
trogen-fixers) or, for cyanobacteria incapable of fixing nitrogen, in
BG-11 medium with 1.5 g l
–1
of NaNO
3
added (Stanier et al. 1971).
The cultivation temperature was 28 C, and the specific conditions
of light and shaking were as given by Johansson and Bergman
(1994).
Detection of auxin-like substances with the Salkowski reagent
Cultures grown for 18 days were centrifuged at 5,000 g for 20 min
at 4 C and then supernatant was purified using a C
18
Sep-Pack
cartridge (Waters, Eschborn, Germany). The culture supernatant
was used for determinations of released, Salkowski-positive indolic
compounds, likely to represent the auxin (Glickmann and Dessaux
1995). The Salkowski reagent, modified according to Pilet and
Chollet (1970), was added to the supernatant in a ratio of 1:1 (v/v).
Concentrations of auxin-like substances were estimated by absor-
bance at 530 nm after 30 min in the dark at room temperature
against a control containing 1 ml of culture medium and 1 ml of
Salkowski reagent.
Extraction and purification
Supernatants and cell extracts of the 10 selected cyanobacteria were
used for detection and estimation of concentrations of IAA in the
cyanobacteria. Cells were harvested by centrifugation at 5,000 g for
20 min at 4 C. The culture supernatant was adjusted to pH 2.8
with 1.0 M HCl and extracted three times with ethyl acetate (1:3, v/
v). Extracts were then evaporated under vacuum at 37 C (Buchi,
Flawil, Switzerland). The aqueous fraction was adjusted to pH 7.0
with 1.0 N NaOH and extracted three times with water-saturated
n-butanol followed by drying in vacuum. The extracts obtained
were filtered through membrane filters (pore size 0.45 lm). Har-
vested cyanobacterial cells were homogenised in liquid nitrogen
using a cold mortar and pestle. The resulting powder was extracted
overnight in 80% methanol containing 10 mg l
–1
butylated hy-
droxytoluene as antioxidant at 4 C. The methanolic fractions were
centrifuged at 4,000 g for 10 min at 4 C, followed by filtering
through glass microfibre filters (pore size 1.2 lm; GF/C, What-
man). The filtrate was then passed through a C
18
Bond Elut solid-
phase extraction column (Varian, Palo Alto, Calif., USA) and
eluted with 80% methanol. Eluates were partitioned with ethyl
acetate and water-saturated butanol as described above. The or-
ganic phases were reduced to dryness with nitrogen and the resi-
dues were re-dissolved in methanol, diluted with 1% acetic acid and
applied to C
18
columns as described above. All extracts were stored
at –18 C and assayed within 48 h.
Enzyme-linked immunosorbent assay (ELISA)
Samples extracted from cells and culture supernatants of cyano-
bacteria were methylated with approximately 1.5 ml of diluted
diazomethane (Cohen 1984) for 5 min at 0 C and dried under
nitrogen. An ELISA test (Phytodetek IAA; Agdia, Elkhart, Ind.,
USA) was used according to the manufacturer’s instructions to
identify IAA. The anti-IAA monoclonal antibody contained in this
kit is documented to show a high specificity for IAA in the
methylated form. The production and characterisation of the high-
affinity monoclonal antibody was as described by Mertens et al.
(1985).
GC-MS analysis
[
13
C
6
]Indole-3-acetic acid (1 ng; 99% enrichment; Cambridge
Isotope Laboratories, Woburn, Mass., USA) as an internal stan-
dard was added before the procedure for extraction and purifi-
cation of IAA from the cyanobacteria mentioned above. Samples
were methylated with diazomethane and dissolved in heptane.
GC-MS analyses were performed according to Edlund et al.
(1995). All samples were processed by the JEOL MS-MP7000 D
data system.
Chlorophyll measurement
The Nostoc cells were collected by centrifugation, the pellets were
ground in liquid nitrogen and the resulting homogenate was dis-
solved in 80% methanol. After 2 h incubation in darkness, cell
debris was removed by centrifugation and the absorbance of the
supernatant was measured against 80% methanol as a blank.
Chlorophyll content (lgml
–1
) was determined from the optical
230
density at 665 nm (OD
665nm
), calculated as OD
665nm
·13.9
(Tandeau de Marsac and Houmard 1988).
Similarity search
Preliminary sequence data of Nostoc punctiforme ATCC 29133 and
Synechococcus WH8102 were obtained from the DOE Joint Ge-
nome Institute (JGI) at http://spider.jgi-psf.org/JGI_microbial/
html ,and Nostoc sp. PCC 7120, Synechocystis sp. PCC 6803 from
the Kazusa DNA Research Institute at http://www.kazusa.or.jp/
cyano/cyano.html. Similarity searches were performed using
BLAST (Altschul et al. 1997) provided at the sites listed above or
included in the Biology Workbench 3.2 package from the San
Diego Supercomputer Center at http://workbench.sdsc.edu. Mul-
tiple sequence alignments were done using CLUSTAL W
(Thompson et al. 1994). The full-length deduced amino
acid sequences of indole-3-pyruvate decarboxylases (IpdCs) from
Azospirillum brazilense, Enterobacter cloacae, Erwinia herbicola and
Pseudomonas putida, tryptophan monooxygenases (IaaMs) and
indole-3-acetamide hydrolases (IaaHs) from Pseudomonas syringae
and Agrobacterium vitis were obtained from the GenBank at the
National Center for Biotechnological Information, NCBI (Gen-
Bank accession numbers P51852, BAA14242, AAB06571,
AAG00523, M11035, M35690, AF061780, AF039169).
DNA analysis
A 1805-bp fragment of the putative ipdC gene from Nostoc sp.
PCC 73102 was PCR-amplified with DynaZyme (Finnzymes,
Espoo, Finland). The primers used were: IPDC-F 5¢-AACGG-
CAAAAGTCACATTCC-3¢ and IPDC-R 5¢-TGAGTTGAC-
GAGGCAGTACG-3¢. The PCR product was ligated into the
Table 1 Screening of release of auxin-like indolic substances from
free-living and symbiotic cyanobacteria using the Salkowski re-
agent. The results are from three independent experiments (after
18 days of cultivation). Symbols represents colour intensity:
+++, strong; ++, average; +, weak; –, no reaction. The true
identity of the isolate denoted Nostoc/Anabaena azollae isolated
from Azolla leaf cavities has been questioned (see Rai 1996)
a
U Unicellular cyanobacteria, NH filamentous non-heterocystous,
FH filamentous heterocystous, N
+
capable of fixing atmospheric
nitrogen, Ex/In extra- or intracellular location of the cyanobacte-
rium in the host
b
Collection Nationale de Cultures de Microoganismes, Institut
Pasteur, Paris (France)
c
Department of Microbiology, University of Helsinki, Helsinki
(Finland)
d
Department of Botany, Stockholm University, Stockholm (Swe-
den)
e
Division of Microbiology, University of Leeds, Leeds (UK)
f
Department of Microbiology, University of California, Davis
(USA)
Cyanobacterial genera Cyanobacterial
species/strain
Morphotype/host plant
a
Auxin-like
Salkowski reaction
Source
Free-living:
Synechocystis PCC 6803 U CNCM
b
Anacystis PCC 6301 U CNCM
b
Gloeothece PCC 6909 U/N
+
+ CNCM
b
Phormidium NH B. Bergman
d
Plectonema PCC 73110 NH/N
+
++ CNCM
b
Symploca PCC 8002 NH/N
+
CNCM
b
Calothrix PCC 7504 FH/N
+
+++ CNCM
b
Calothrix PCC 7103 FH/N
+
+++ CNCM
b
Scytonema PCC 7110 FH/N
+
CNCM
b
Chlorogloeopsis PCC 6912 FH/N
+
CNCM
b
Fischerella PCC 7521 FH/N
+
CNCM
b
Nostoc 268 FH/N
+
+ T. Vaara
c
Nostoc/Anabaena PCC 7120 FH/N
+
CNCM
b
Nostoc PCC 6720 FH/N
+
CNCM
b
Nostoc PCC 6310 FH/N
+
+ CNCM
b
Nostoc PCC 7107 FH/N
+
CNCM
b
Symbiotic/N
+
:
Nostoc Pc Lichen (Ex) +++ B. Bergman
d
Nostoc LBG1 Liverwort (Ex) + D. Adams
e
Nostoc PCC 9305 Liverwort (Ex) + CNCM
b
Nostoc 7801 Hornwort (Ex) + J.C. Meeks
f
Nostoc/Anabaena azollae Newton’s isolate Water-fern (Ex) +++ J.C. Meeks
f
Nostoc 8071 Cycad (Ex) + J.C. Meeks
f
Nostoc MacS Cycad (Ex) + CNCM
b
Nostoc PCC 7422 Cycad (Ex) CNCM
b
Nostoc PCC 73102 Cycad (Ex) ++ CNCM
b
Nostoc 8005 Gunnera (In) J.C. Meeks
f
Nostoc Chr 891 Gunnera (In) + C. Johansson
d
Nostoc Chr 892 Gunnera (In) ++ C. Johansson
d
Nostoc Chr 893 Gunnera (In) + C. Johansson
d
Nostoc Chr 894 Gunnera (In) +++ C. Johansson
d
Nostoc 7901 Gunnera (In) + J.C. Meeks
f
Nostoc 8001 Gunnera (In) ++ J.C. Meeks
f
Nostoc PCC 9229 Gunnera (In) ++ CNCM
b
Nostoc PCC 9231 Gunnera (In) CNCM
b
231
pCR 2.1 vector and cloned in Escherichia coli TOP 10F¢ (Invitro-
gen, Groningen, The Netherlands). The insert obtained was used as
a probe in Southern analysis. Total DNA from 11 cyanobacterial
strains (Gloeothece sp. PCC 6909, Plectonema sp. PCC 73110,
Nostoc sp. PCC 6720, Nostoc sp. PCC 7120, Nostoc sp. 268, Nostoc
sp. Pc, Nostoc sp. LBG1, Nostoc sp. PCC 73102, Nostoc sp.
Chr 894, Nostoc sp. PCC 9229 and Nostoc/Anabaena azollae) was
isolated as in Cohen et al. (1994). For each strain, total DNA (8–
10 mg) was digested with EcoRI (Amersham Pharmacia, Uppsala,
Sweden) and separated on a 1% agarose gel with 1·TAE buffer
(Sambrook et al. 1989). The DNA was transferred onto Hybond
N+ membrane (Amersham Pharmacia). Southern hybridisation
was performed with the ECL direct nucleic acid labelling and de-
tection system (Amersham Pharmacia) according to the manufac-
turer’s instructions.
Results
Screening for the release of auxin-like indolic
compounds
As a first step, 34 free-living and symbiotically compe-
tent cyanobacterial strains were screened for their ability
to produce auxin-like substances (Table 1). Represen-
tatives from all major cyanobacterial morphotypes were
analysed: unicellular (U), filamentous non-heterocystous
(NH) and filamentous heterocystous (FH; heterocysts
are cells that fix atmospheric nitrogen). Moreover, the
selection embraced cyanobacterial strains with known
symbiotic competence, originally isolated from five host-
plant groups, from a lichenized fungus to the angio-
sperm, Gunnera.
The auxin-like substances were determined using the
Salkowski reagent. As this reagent represents a simple
and fast assay, and as it is known to recognise indolic
compounds such as IAA (Glickmann and Dessaux 1995),
it was used for the determination of IAA by Hartman
et al. (1983) and Zimmer et al. (1991). Twenty-one of the
34 cyanobacterial strains tested (approx. 62%) showed a
positive colour reaction indicative of a release of an IAA-
like substance into the culture supernatant (Table 1). Of
the 16 free-living strains, about 38% released IAA-like
substances. Positive representatives were found among
all three cyanobacterial morphotypes (U, NH, FH;
Table 1), and the nitrogen-fixing heterocystous genus
Calothrix (PCC 7103 and PCC 7504) appeared particu-
larly potent. No correlation between release of IAA-like
substances and morphological complexity was detected.
The most complex representatives (Scytonema, Chloro-
gloeopsis and Fischerella), belonging to group-V cyano-
bacteria sensu Rippka et al. (1979), showed no
Salkowski-positive response. Neither was there any cor-
relation between the Salkowski response and nitrogen-
fixation capacities (Table 1).
Of the 18 symbiotic Nostoc strains (all belonging to
group-IV cyanobacteria), 83% released IAA-like sub-
stances. The highest levels of release were seen in Nostoc
isolated from the lichenized fungus Peltigera, from the
angiosperm Gunnera and the putative isolate from the
water-fern Azolla, but all host divisions were represented
among Salkowski-positive strains. There was no corre-
lation between release of IAA-like substances and
whether the cyanobacterium forms an intra- or extra-
cellular (In/Ex) symbiosis (Table 1).
Immunological identification of IAA
To obtain a more accurate qualitative identification and
quantitative determinations of the IAA released as well
as of the cellular IAA contents, an immunological ap-
proach was used. Ten of the screened cyanobacteria
(Table 1) were selected for examination using a specific
monoclonal anti-IAA antibody using ELISA (Weiler
et al. 1981). This assay detects the IAA methyl ester
down to 10–20 fmol, and is claimed to be as sensitive as
GC-MS (Pengelly et al. 1981). As is obvious from
Table 2, an IAA-positive response was obtained from
the strains examined, although the quantities varied,
with the unicellular Gloeothece PCC 6909 accumulating
and releasing the least. On average, both the endogenous
and exogenous IAA levels were of the same order of
magnitude within the two groups (Table 2). In general,
Table 2 Endogenous and
exogenous levels of IAA
produced by cyanobacteria
grown in the presence and
absence of tryptophan (Trp) as
evidenced by an ELISA test.
The tests were performed after
2 weeks of cultivation. The
values represent the mean ±
SD from two independent ex-
periments with three replicates
each
Cyanobacterial strain
a
Endogenous IAA Exogenous IAA
(pmol mg
–1
Chl a) (pmol mg
–1
Chl a)
–Trp +Trp –Trp +Trp
Free-living:
Gleoethece 3.5±0.6 3.3±0.3 1.2±0.5 1.2±0.6
Plectonema 12.8±0.5 14.9±0.2 12.6±0.6 10.5±0.5
Nostoc 268 10.1±0.4 30.9±0.8 5.2±0.3 9.1±0.4
Nostoc 7120 15.3±0.8 21.1±0.5 9.9±0.6 12.6±0.4
Nostoc 6720 6.0±0.4 20.8±1.1 11.3±0.7 19.5±0.3
Symbiotic:
Nostoc Pc 16.2±0.5 26.8±0.7 23.5±0.9 25.7±1.2
Nostoc/Anabena 12.5±0.5 20.7±1.4 10.1±0.8 18.6±1.1
Nostoc 73102 24.2±1.1 42.2±1.5 19.9±0.7 34.0±1.2
Nostoc Chr 894 7.2±0.6 37.6±1.6 15.2±0.8 24.7±0.9
Nostoc 9229 14.9±0.6 28.5±0.7 8.2±0.3 23.1±0.6
a
For morphotype, strain identification and source of the cyanobacteria used, see Table 1
232
the symbiotic isolates showed higher endogenous IAA
levels and released more IAA than the free-living
counterparts.
Tryptophan (Trp) is considered as the main precursor
for the biosynthesis of IAA in plants and microorgan-
isms, but with several possible pathways and interme-
diates involved in generating the final product, IAA.
Adding 500 mg l
–1
Trp to the cyanobacterial cultures
stimulated the endogenous accumulation of IAA about
2-fold in both free-living and symbiotic strains, while
IAA release was less enhanced by Trp (Table 2). Nostoc
PCC 73102 accumulated about 42 pmol and released
34 pmol IAA per mg chlorophyll a (Chl a) after 2 weeks
in the presence of Trp, the highest quantities noted.
Accumulation and release of IAA
in two Nostoc strains
Two of the Nostoc strains were selected for comparative
analyses of their IAA production during a 3-week
growth cycle. The cyanobacteria were the free-living
Nostoc 268, originally isolated from the Baltic Sea and
not known to be involved in any symbiotic interactions
and incapable of infecting Gunnera (Johansson and
Bergman 1994), and the symbiotically competent Nos-
toc 9229. The latter readily infects Gunnera cells intra-
cellularly. Both Nostoc strains continuously
accumulated IAA endogenously and released IAA dur-
ing the 3-week growth period, and the IAA concentra-
tions in the free-living Nostoc 268 were in general lower
than in the symbiotic isolate (Fig. 1a, b, Table 2).
However, IAA was not detected after 24 h of growth,
neither was a release seen in the absence of Trp within
72 h. Maximum endogenous IAA levels for Nostoc 9229
were about 20 pmol IAA per mg Chl a and that of IAA
release about 15 pmol per mg Chl a (Fig. 1a, b). Addi-
tion of Trp again stimulated IAA levels, the endogenous
levels to a greater extent than the release (Fig. 2a, b),
and those of the symbiotic Nostoc more than those of
the free-living strain.
Chemical identification of IAA by GC-MS
To unequivocally demonstrate the chemical authenticity
of IAA in cyanobacteria, GC-MS was used. Represen-
tative mass chromatograms of two samples for the two
cyanobacteria are shown in Fig. 3. The true identity of
IAA in the cell extracts was hereby verified. This was
also demonstrated for eight additional samples repre-
senting the two Nostoc strains, analysed at various time
points and under different growth conditions (data not
shown).
A comparison of IAA concentrations obtained using
the ELISA immunoassay and the chemical GC-MS an-
alyses is shown in Table 3. A high positive correlation
between the two methods was observed. However, the
GC-MS method consistently demonstrated higher
endogenous as well as exogenous IAA concentrations,
irrespective of Nostoc strain examined or time expired.
As previously, addition of Trp stimulated both accu-
mulation and release of IAA in the two strains and this
was particularly evident using GC-MS. The latter assay
revealed about 2-fold higher IAA quantities than the
ELISA test (Table 3). After 2 weeks in the presence of
Trp, the endogenous levels of IAA amounted to 67 and
52 pmol per mg Chl a for the free-living and the sym-
biotic strains, respectively.
Identification of genes with homology
to IAA biosynthetic genes
A number of bacterial genes involved in the biosynthesis
of IAA have been identified and characterised. This to-
gether with the availability of several complete cyano-
bacterial genomes provoked us to search for genes
encoding homologues of bacterial and plant proteins
involved in IAA synthesis. However, a similarity search
for genes encoding enzymes of the major indole-3-
acetamide pathway, tryptophan 2-monooxygenases
(iaaM) and indole-3-acetamide hydrolases (iaaH; Patten
and Glick 1996) in Pseudomonas syringae and Agro-
bacterium vitis (GenBank accession numbers M11035,
Fig. 1 Accumulation (a) and release (b) of IAA by a free-living and
a symbiotically competent Nostoc, as evidenced by ELISA. The
values represent the mean (±SD) from two independent experi-
ments with three replicates each
233
M35690, AF061780, AF039169), gave no positive result.
This was the case in the three cyanobacterial genomes
tested: the unicellular Synechocysis sp. PCC 6803, the
free-living filamentous Nostoc/Anabaena sp. PCC 7120
and the symbiotically competent Nostoc punctiforme
ATCC 29133 (=Nostoc PCC 73102). Similarly, nega-
tive results were obtained for the P. syringae gene
encoding IAA-lysine synthetase (iaaL), converting IAA
to a less bioactive form (Roberto et al. 1990; Glickmann
et al. 1998). However, the genome of the IAA-releasing
N. punctiforme ATCC 29133 (Nostoc PCC 73102; see
Tables 1, 2) contains a gene coding for a predicted thi-
amine pyrophosphate-requiring enzyme, TPP-enzyme.
This 558-amino-acid-long protein shows similarities to
indole-3-pyruvate decarboxylases, encoded by the ipdC
gene, from Erwinia herbicola (GenBank accession num-
ber AAB06571), Enterobacter cloacae (BAA14242) and
P. putida (AAG00523). IpdC catalyses the conversion of
indole-3-pyruvate to indole-3-acetaldehyde (Koga 1995).
The amino acid identity values ranged from 41% to 43%
between IpdC from these bacteria and the homologue in
N. punctiforme (Nostoc PCC 73102). Comparison of
deduced amino acid sequences of IpdC of Enterobacter
cloacae with that of Azospirillum brasilense gave a 43%
homology (Costacurta et al. 1994). A full-length de-
duced amino acid sequence alignment of the putative
cyanobacterial and the bacterial IpdCs is presented in
Fig. 4. The alignment convincingly demonstrates that
the degree of homology between the predicted TPP-en-
zyme from N. punctiforme and IpdC from Enterobacter
cloacae, Erwinia herbicola and P. putida is higher than
the degree of homology between these bacteria and IpdC
from A. brasilense. Moreover, all amino acid residues
involved in the formation of the active site of IpdCs
(Koga 1995) are found in the conserved positions in the
sequence from Nostoc. This suggests that the cyano-
bacterial TPP-enzyme is contained within the bacterial
radiation of IpdCs. BLAST searches for the homologues
of this TPP-enzyme in public databases gave best scores
to the known and putative IpdCs (data not shown). In
addition, the genome of N. punctiforme contains several
copies of the aspartate aminotransferases (EC 2.6.1.1.)
Fig. 2 Accumulation (a) and release (b) of IAA by a free-living and
a symbiotically competent Nostoc in the presence of the putative
IAA precursor tryptophan (500 mg l
–1
), as evidenced by ELISA.
The values represent the mean (±SD) of two independent
experiments with three replicates each
Fig. 3 GC-MS identification of
endogenous IAA in the free-
living Nostoc 268 (in absence of
tryptophan; a, b) and in the
symbiotic isolate Nostoc
PCC 9229 (in presence of
tryptophan; c, d), both assayed
after 2 weeks of cultivation.
The peaks at m/z 202.1050
represent the IAA-methyl trim-
ethylsilyl ester (a, c) and the
peaks at m/z 208.1250 are indi-
cative of [
13
C
6
]IAA-methyl
trimethylsilyl ester (b, d), where
the [
13
C
6
]IAA was used as
internal standard (IS)
234
required for the first step of this pathway converting
tryptophan into indole-3-pyruvate. In the genome of
Synechocystis sp. PCC 6803, the product of merR shows
a high degree of similarity to arylacetonitrilase from
Alcaligenes faecalis (34% identity) and nitrilase II from
Arabidopsis thaliana (36% identity), enzymes involved in
the indole-3-acetonitrile pathway of IAA synthesis
(Bartling et al. 1994; Kobayashi et al. 1993). A similar
open reading frame (ORF) was also found in the ge-
nome of unicellular Synechococcus WH8102 (data not
shown).
Presence of the ipdC homologue in cyanobacteria
To complement our similarity search for putative IAA-
biosynthesis enzymes using BLAST, Southern hybridi-
sation of total DNA from 11 strains of cyanobacteria
(ranging from unicellular to filamentous heterocystous;
see Table 1) was performed using the PCR amplified and
cloned fragment of the putative ipdC gene from Nostoc
PCC 73102. Four positive ipdC signals were detected in
DNA of the seven strains shown in Fig. 5: in Nostoc 268,
Nostoc PCC 73102, Nostoc PCC 9229, and Anabaena/
Nostoc azollae. Four additional strains were also tested:
Gloethece sp. PCC 6909, Plectonema sp. PCC 73110,
Nostoc/Anabaena PCC 7120 and Nostoc PCC 6720, but
no positive ipdC response was obtained (data not
shown). Hence, it is apparent that although the majority
of these cyanobacteria responded positively in the IAA-
ELISA test (Nostoc LBG1 was not tested; Table 2), only
4 of 11 showed a response to the heterologous ipdC
probe.
Discussion
Here we provide the first evidence for the capacity
among cyanobacteria to accumulate IAA endogenously
and to release the hormone. We also demonstrate the
presence of homologues of the ipdC gene in some, but
not in all of the strains tested, a gene of key importance
in IAA biosynthesis. First, the Salkowski dye showed
release of auxin-like substances in 21 of the 34 cyano-
bacterial strains screened. The authenticity of IAA in
cell extracts and culture supernatants was then verified
using the specific monoclonal IAA antibody on 10 of the
screened Nostoc strains, and finally the chemical identity
of the IAA accumulated and released by 2 of these
strains was demonstrated using GC-MS.
It is apparent that the Salkowski reagent is the least
sensitive of the three methods used as some of the Sal-
kowski-negative strains responded positively using EL-
ISA and GC-MS (compare Tables 1 and 2), while the
reverse was never the case. Moreover, there is a positive
correlation between the ELISA and the GC-MS IAA
levels. But the GC-MS analyses in general gave higher
IAA levels than did the IAA antibody–ELISA test, and
particularly so in the presence of Trp. The reason for
this discrepancy is not known but the GC-MS method is
considered to be the most accurate.
The IAA levels in cyanobacterial suspensions are
difficult to compare with those reported earlier for
bacteria and other photosynthetic organisms, such as
plants, cyanobacteria being strikingly dissimilar in their
appearance. For instance, the IAA production for bac-
teria is often expressed per number of cells or per ml
culture medium. The cyanobacterial cell volumes are
about two orders of magnitude larger than those of most
bacteria, which leads to different packing efficiencies,
and makes comparisons of IAA levels per number of
cells or per ml unsuitable. However, cyanobacterial
cultures and protoplast suspensions are quite compara-
ble in their characters. After 2 weeks of growth in the
presence of Trp, the GC-MS recorded endogenous IAA
levels for the two Nostoc strains, amounted to about 12
and 9 pg IAA per lg Chl a, respectively (Table 3). These
levels are comparable to the about 146 pg IAA per lg
Chl a recorded after 7 weeks for a tobacco protoplast
suspension (Sitbon et al. 2000).
Our data also indicate that cyanobacteria may be
able to convert exogenous Trp into IAA (Fig. 2,
Tables 2, 3). However, it is possible that Trp was used as
a source of nitrogen during the initial stages of cultiva-
tion as the increase in IAA concentrations due to Trp
was most obvious during the later stages of cultivation.
Table 3 Comparative analyses
of IAA levels in a free-living
and a symbiotic Nostoc strain in
the presence and absence of
Trp. Data obtained by ELISA
and by GC-MS. The values
represent the mean ± SD from
two independent experiments
with three replicates each. a
IAA accumulation, r IAA
release, n.d. IAA not detected
Cyanobacterial
strain
IAA a/r Trp added (+) Cultivation period IAA (pmol mg
–1
Chl a)
ELISA GC-MS
Nostoc 268 a 72 h 10.0±0.5 15.8
Nostoc 268 a 2 weeks 10.1±0.4 12.5
a
Nostoc 268 a + 2 weeks 30.9±0.8 67.0
Nostoc 268 r 72 h n.d. 0.3
Nostoc 268 r + 1 weeks 4.0±0.4 11.3
Nostoc 9229 a 72 h 5.1±0.6 8.9
Nostoc 9229 a 2 weeks 14.9±0.6 17.8
Nostoc 9229 a + 2 weeks 28.5±0.7 52.0
b
Nostoc 9229 r 72 h n.d. 0.7
Nostoc 9229 r + 1 weeks 14.1±0.3 31.8
a
From Fig. 3a, b
b
From Fig. 3c, d
235
As Trp inhibits both heterocyst development and
nitrogen fixation, while its analogue, 7-azatryptophan,
provokes heterocyst differentiation in free-living An-
abaena cylindrica (Pain et al. 2000, p. 147), it is of in-
terest to examine the role of Trp on cyanobacterial
performance in general.
The above-mentioned observations related to IAA
biosynthesis, together with the results of the gene/amino
acid similarity BLAST search, suggest that a major
pathway for the production of IAA, at least in some
cyanobacteria, may be via the indole-3-pyruvic acid
pathway. The key enzyme (indole-3-pyruvate decar-
boxylase) of this pathway is encoded by ipdC and our
analyses revealed that ipdC-like genes are present in
some Nostoc species. Recently, it has been shown that in
A. brasilense, IpdC may account for most of the IAA
produced, that IAA is synthesised in a cell-density de-
pendent manner, and that accumulated IAA acts as an
autoinducer of IAA biosynthesis (Van de Broek et al.
1999). A similar IAA-accumulation effect was noted in
older cultures of cyanobacteria (Figs. 1, 2). Comparative
analysis of the amino acid sequences of several bacterial
IpdC enzymes led us to conclude that the TPP-enzyme in
Nostoc may be IpdC. This, however, will now need to be
Fig. 4 Full-length amino acid sequence alignment of deduced
indole-3-pyruvate decarboxylases (IpdCs) from IAA-producing
bacteria and a predicted TPP-enzyme from the cyanobacterium
Nostoc punctiforme ATCC 29133 (PCC 73102). Bacteria included
were Azosprillum brasilense, Enterobacter cloacae, Erwinia herbicola
and Pseudomonas putida (GenBank accession numbers P51852,
BAA14242, AAB06571, AAG00523). Identical amino acids are
shown in red, conserved in blue and similar amino acids in green.
The active site is marked by a cross, conserved residues in the
vicinity of the active site by bold arrows, the putative Mg
2+
and
diphosphate-binding domain by a bold line, and W denotes the
conserved tryptophan residue (after Koga 1995)
Fig. 5 Southern-blot analysis of total DNA from seven Nostoc
strains probed with a fragment of the putative ipdC gene from
Nostoc PCC 73102. Lanes: 1, Nostoc 268; 2, Nostoc LBG1; 3,
Nostoc PCC 73102; 4, Nostoc PCC 9229; 5, Nostoc Pc; 6, Nostoc-
Chr 894; 7, Anabaena/Nostoc azollae. All except Nostoc 268 (f) are
symbiotic isolates (s). DNA markers are given to the left
236
verified, as will the existence of other possible pathways
for IAA production in cyanobacteria.
Cyanobacteria may respond developmentally to
changed phytohormone signals both when free-living
and when in planta. Cyanobacteria are morphologically
more complex than most prokaryotes and some are
able to form several cell types (Rippka et al. 1979; Fay
1992). Most of these cell types appear in a strict timely
fashion in compatible Nostoc spp. infecting Gunnera.
First, small-celled motile hormogonia develop and la-
ter, after host cell penetration, these revert back into
multi-heterocystous vegetative filaments (Adams 2000;
Rai et al. 2000). As it has also been shown that IAA
triggers differentiation of cyanobacterial hormogonia
(Bunt 1961), the effect of IAA (and biosynthetic pre-
cursors) on cyanobacterial development in general is
now being tested.
Symbiotic cyanobacteria, like other plant-interacting
prokaryotes, whether beneficial (Azospirillum and
Bradyrhizobium) or pathogenic (Erwinia and Agrobac-
terium), may use the capacity to produce and release this
phytohormone as a strategy to influence endogenous
plant hormone levels and thereby plant performance
(Glick et al. 1999). Several developmental events are
known to be triggered in both partners during estab-
lishment of cyanobacterial–plant symbiosis. During
infection of the mucilage-secreting stem glands of Gun-
nera (Bergman et al. 1992, 1996; Rasmussen et al. 1994),
compatible Nostoc strains stimulate mitotic activity in
host cells close to the site of cyanobacterial host cell
penetration (Johansson and Bergman 1992). Also,
cyanobacteria provoke growth of plant protrusions
ramifying through the mucilage-filled cavities infected
by cyanobacteria in bryophytes and Azolla (Rodgers
and Stewart 1977; Peters 1991). IAA is known to cause
cell wall changes in plants, to enhance cell permeability,
to cause swelling of exposed plant cells and to stimulate
the production of a second hormone, ethylene (Liu et al.
1982). It is now of interest to examine whether cyano-
bacteria also make use of their IAA-release to initiate
the development of the symbiotic plant tissues.
In conclusion, our data for the first time reveal that
free-living and, in particular, symbiotic cyanobacteria,
are able to accumulate and release the phytohormone
indole-3-acetic acid, and that exogenous Trp stimulates
these processes. The next target in our investigation is to
isolate the putative ipdC gene from cyanobacteria and to
follow its expression, as well as to elucidate IAA-bio-
synthesis pathways and the role of IAA in cyanobacte-
rial cell differentiation in and plant symbiosis.
Cyanobacteria with their complex life cycle and their
unique ability to simultaneously fix nitrogen and per-
form oxygenic photosynthesis appear to share the same
mechanism for accumulation of IAA with some other
plant-interacting bacteria as well as with plants. In view
of the fact that the progenitors of extant cyanobacteria
were ancestors of plastids, the possible evolutionary link
between cyanobacterial biosynthesis of IAA and that of
plants is another aspect worth examining.
Acknowledgements The Department of Forest Genetics and Plant
Physiology, Sweden University of Agricultural Sciences (Umea
˚
,
Sweden) is gratefully acknowledged for GC-MS analyses of IAA.
We also thank the Royal Swedish Academy of Sciences, the
Swedish Natural Science Research Council and the Forestry and
Agricultural Research Council for financial support.
References
Adams DG (2000) Symbiotic interactions. In: Whitton BA, Potts
M (eds) The ecology of cyanobacteria, Kluwer Academic
Publishers, Dordrecht, The Netherlands, pp 523-561
Altschul SF, Madden TL, Scha
¨
ffer AA, Zhang J, Zhang Z, Miller
W, Lipman DJ (1997) Gapped BLAST and PSI-BLAST: a new
generation of protein database search programs. Nucleic Acids
Res 25:3389–3402
Bartling D, Seedorf M, Schmidt RC, Weiler EW (1994) Molecular
characterization of two cloned nitrilases from Arabidopsis tha-
liana: key enzymes in accumulation of the plant hormone
indole-3-acetic acid. Proc Natl Acad Sci 91:6021–6025
Bergman B, Johansson C, So
¨
derba
¨
ck E (1992) The NostocGun-
nera symbiosis. New Phytol 122:379–400
Bergman B, Matveyev A, Rasmussen U (1996) Chemical signalling
in cyanobacterial-plant symbioses. Trends Plant Sci 1:191–197
Bunt JS (1961) Isolation of bacteria-free cultures from hormogone-
producing blue-green algae. Nature 192:1275–1276
Cohen JD (1984) Convenient apparatus for the generation of small
amounts of diazomethane. J Chromatogr 303:193–196
Cohen MF, Wallis JG, Campbell EL, Meeks JC (1994) Transposon
mutagenesis of Nostoc sp. strain ATCC 29133, a filamentous
cyanobacterium with multiple cellular differentiation alterna-
tives. Microbiol 140: 3233–3240
Costacurta A, Vanderleyden J (1995) Accumulation of phytohor-
mones by plant-associated bacteria. Crit Rev Microbiol 21:1–18
Costacurta A, Keijers V, Vanderleyden J (1994) Molecular
cloning and sequence analysis of an Azospirillum brasilense
indole-3-pyruvate decarboxylase gene. Mol Gen Genet
243:463–472
Christiansen-Weniger C (1998) Endophytic establishment of dia-
zotrophic bacteria in auxin-induced tumors of cereal crops. Crit
Rev Plant Sci 17:55–76
Douglas SE (1994) Chloroplast origins and evolution. In: Bryant
DA (ed) The molecular biology of cyanobacteria. Kluwer Ac-
ademic Publishers, Dordrecht, The Netherlands, pp 91–118
Edlund A, Eklof S, Sundberg B, Moritz T, Sandberg G (1995) A
microscale technique for gas chromatography–mass spectrom-
etry measurements of picogram amounts of indole-3-acetic acid
in plant tissues. Plant Physiol 108:1043–1047
Glick BR, Patten CL, Holguim G, Penrose DM (1999) Biochemical
and genetic mechanisms used by plant growth promoting bac-
teria. ICP, Covent Garden, London
Glickmann E, Dessaux Y (1995) A critical examination of the
specificity of the Salkowski reagent for indolic compounds
produced by phytopathogenic bacteria. Appl Environ Micro-
biol 61:793–796
Glickmann E, Gardan L, Jacquet S, Hussain S, Elasri M, Petit A,
Dessaux Y (1998) Auxin production is a common feature of
most pathovars of Pseudomonas syringae. Mol Plant-Microbe
Interact 11:156–162
Fay P (1992) Oxygen relations of nitrogen fixation in cyanobacte-
ria. Microbiol Rev 56:340–373
Hartman A, Singh M, Klingmuller W (1983) Isolation and char-
acterization of Azospirillum mutants excreting high amounts of
indole-acetic acid. Can J Microbiol 29:916–923
Johansson C, Bergman B (1992) Early events during the estab-
lishment of GunneraNostoc symbiosis. Planta 188:403–413
Johansson C, Bergman B (1994) Reconstitution of the symbiosis of
Gunnera manicata Linden: cyanobacterial specificity. New
Phytol 126:643–652
Kobayashi M, Izui H, Nagasawa T, Yamada H (1993) Nitrilase in
accumulation of the plant hormone indole-3-acetic acid from
237
indole-3-acetonitrile: cloning of the Alcaligenes gene and site-
directed mutagenesis of cysteine residues. Proc Natl Acad Sci
90:247–251
Koga J (1995) Structure and function of indolepyruvate decar-
boxylase, a key enzyme in indole-3-acetic acid accumulation.
Biochim Biophys Acta 1249:1–13
Liu ST, Perry KL, Schardl CL, Kado CL (1982) Agrobacterium Ti
plasmid indoleacetic acid gene is required for crown gall on-
cogenesis. Proc Natl Acad Sci 79:2812–2816
Mertens R, Eberle J, Arnscheidt A, Ledebur A, Weiler EW (1985)
Monoclonal antibodies to plant regulators. II. Indole-3-acetic
acid. Planta 166:389–393
Pain R, Duggan PS, Adams DG (2000) Creation of a tryptophan
auxotroph to study the role of tryptophan in heterocyst devel-
opment. In: Abstracts of the 10th international symposium on
phototrophic prokaryotes, Barcelona, Spain, August 26–31
2000, p 147
Patten CL, Glick BR (1996) Bacterial accumulation of indole-3-
acetic acid. Can J Microbiol 42:207–220
Pengelly WL, Bandurski RS, Schulze A (1981) Validation of a
radioimmunoassay for indole-3-acetic acid using gas chroma-
tography–selected ion monitoring mass spectrometry. Plant
Physiol 68:96–98
Peters GA (1991) Azolla and other plant–cyanobacteria symbioses:
aspects of form and function. Plant Soil 137:193–210
Pilet PE, Chollet R (1970) Sur le dosage colorimetrique de l’acide
indolylacetique. C R Acad Sci Ser D 271:1675–1678
Rai AN (1996) Handbook of symbiotic cyanobacteria, CRC Press,
Boca Raton, USA
Rai AN, So
¨
derba
¨
ck E, Bergman B (2000) Cyanobacterium–plant
symbioses. New Phytol 147:449–481
Rasmussen U, Johansson C, Bergman B (1994) Early communi-
cation in the GunneraNostoc symbiosis: plant-induced cell
differentiation and protein accumulation in the cyanobacteri-
um. Mol Plant Microbe Interact 7:696–702
Rippka R, Deruells J, Waterbury JB, Herdman M., Stanier R
(1979) Generic assignment, strain histories and properties of
pure cultures of cyanobacteria. J Gen Microbiol 111:1–61
Roberto FF, Kle HJ, White F, Nordeen R, Kosuge T (1990)
Expression and fine structure of the gene encoding
N-epsilon-(indole-3-acetyl)-
L-lysine synthetase from Pseudo-
monas savastanoi. Proc Natl Acad Sci 87:5797–5801
Rodgers GA, Stewart WDP (1977) The cyanophyte–hepatic
symbiosis. Morphology and physiology. New Phytol 78:441–
458
Sambrook J, Fritsch EF Maniatis T (1989) Molecular cloning: a
laboratory manual, 2nd edn. Cold Spring Harbor Laboratory
Press, Cold Spring Harbor, NY
Schopf JW (2000) The fossil record: tracing the roots of the cy-
anobacterial lineage. In: Whitton BA, Potts M (eds) The ecol-
ogy of cyanobacteria. Kluwer Academic Publishers, Dordrecht,
The Netherlands, pp 13–35
Sitbon F, A
˚
stot C, Edlund A, Crozier A, Sandberg G (2000) The
relative importance of tryptophan-dependent and tryptophan-
independent biosynthesis of indole-3-acetic acid in tobacco
during vegetative growth. Planta 211:715–721
Stanier RY, Kunisawa R, Mandel M, Cohen-Bazire G (1971) Pu-
rification and properties of unicellular blue-green algae (order
Chroococcales). Bacteriol Rev 35:171–205
Tandeau de Marsac N, Houmard J (1988) Complementary chro-
matic adaptation: physiological conditions and action spectra.
Methods Enzymol 167:318–328
Thompson JD, Higgins DG, Gibson TJ (1994) CLUSTAL W:
improving the sensitivity of progressive multiple sequence
alignment through sequence weighting, position-specific gap
penalties and weight matrix choice. Nucleic Acids Res 22:4673–
4680
Van de Broek A, Lambrecht M, Eggermont K, Vanderleyden J
(1999) Auxins upregulate expression of the indole-3-pyruvate
decarboxylase gene in Azospirillum brazilense. J Bacteriol
181:1338–1342
Weiler EW, Jourdan PS, Conrad W (1981) Levels of indole-3-acetic
acid in intact and decapitated coleoptiles as determined by a
specific and highly sensitive solid-phase enzyme immunoassay.
Planta 153:561–571
Zimmer V, Aparicio C, Elmerich C (1991) Relationship between
tryptophan accumulation and indole-3-acetic acid production
in Azospirillum: identification and sequencing of a TrpGDC
cluster. Mol Gen Genet 29:41–51
238
... Another mechanism is used to convert indole-3-acetic aldehyde from tryptophan, resulting in the formation of an intermediate tryptamine (Pollmann et al., 2009). Most Cyanobacteria biosynthesize IAA by converting tryptophan to indole-3-acetonitrile (Sergeeva et al., 2002). ...
Chapter
The majority of microbial communities on plants are composed of prokaryotes, particularly bacterial domain members. These plant-associated bacteria can help plants improve their growth by enhancing nutrient uptake and increasing plant resistance to various diseases and environmental stresses. It has been discovered that the soil contains a large number of bacteria, often in the range of 10^8-10^9 bacteria per gram of soil. However, only about 1% of these bacteria can be cultured. In soils that have been exposed to environmental stress, the population of culturable bacteria may be as low as 10^4 cells per gram of soil.
... This may be due to more release of carbohydrate/sugar rich exudates or IAA in media, which may influence the growth of microbial biomass in soil as indirectly reflected by increased soil dehydrogenase activity in soil. Cyanobacterial cultures are known to excrete a diverse range of metabolites, including sugars, amino acids, which can stimulate growth, not only in soil, but also in soil -less media and hydroponics (Misra and Kaushik 1989;Sergeeva et al. 2002;Karthikeyan et al. 2009;Bharti et al. 2020), including those used in the present investigation. ...
Article
Full-text available
The nutrient-enriching properties of cyanobacteria were explored to enhance growth and fruit quality in a popular tomato variety Pusa Cherry tomato 1. The dried biomass of plant growth promoting and nitrogen-fixing cyanobacterium Anabaena laxa (A. laxa) and biofilm (Anabaena torulosa-Trichoderma viride; An-Tr) was used to prime the nursery potting mixes. An-Tr significantly enhanced pigment content and root volume of the nursery-raised seedlings by 27 and 50% respectively. Transplantation after 1-month into beds in a climate regulated polyhouse was done and plant, soil samples analyzed. Sampling after 8 WAT (Weeks after transplanting) illustrated the better performance of An-Tr with 30% increase over control, in terms of leaf pigments. This was followed by soil drenching using formulations and foliar application with water/Fe-EDTA, at weekly intervals, from pre-flowering stage (8 WAT) onwards. An-Tr soil drench + water foliar treatment led to significantly superior quality of fruits, when analyzed at first harvest (21 WAT). Ascorbic acid in fruits reported a significant positive correlation with lycopene content (0.83), fruit fresh weight, soil urease activity (0.71. 0.72, respectively), as also with root volume (0.94) and chlorophyll content (0.9) of nursery potting media. Fruit yield was significantly higher, by 20–30% over control, in treatments receiving soil drench of An-Tr + water as foliar or in A. laxa; water drench + Fe-EDTA as foliar. A promising nutri-fertigation strategy was developed, using cyanobacteria to prime nursery, followed by use as drench, along with water/Fe-EDTA as foliar spray, from flowering stage onwards, to improve growth, yield, and quality of cherry tomato.
Article
Full-text available
Rice microbes ecology
Chapter
Photosynthetic organisms grow in diverse habitats and also continuously deal with the fluctuating environmental conditions. The changes in the environmental factors such as intensity and quality of light, salt and nutrient concentrations, temperature extremes, and heavy metal contaminations widely affect the growth and photosynthetic efficiency of the phototrophs. However, these organisms display outstanding ability to cope up with such adverse conditions present on the earth, which is the result of resilience in the photosynthesis. Phototrophs show resistance against ever changing environmental conditions up to a certain level by maintaining their photosynthetic yield. Some phytochemicals such as secondary metabolites and phytohormones play crucial role in the resilient behavior of photosynthesis. Among phytometabolites, alpha tocopherol plays a potent role in scavenging reactive oxygen species under adverse conditions. In phototrophs, during the synthesis of alpha-tocopherol, gamma tocopherol, a phytometabolite is converted into alpha tocopherol via gamma tocopherol methyl transferase (gamma TMT) enzyme in final rate-limiting step. This enzyme has been successfully characterized in plants, but annotation and characterization are still lacking in cyanobacteria. So, diversity and phylogenetic analyses of methyltransferase domain, responsible for gamma TMT activity, have been done in this chapter. After bioinformatic analysis, it was found that some amino acid residues are conserved among domain found in all three groups of phototrophs, and these residues may be crucial for the activity of gamma TMT. Additionally, molecular docking was performed between gamma TMT-containing domain protein (Nostoc sphaeroides CCNUC1) and gamma-tocopherol to validate the interaction between them. Moreover, there are some literature that individually define the role of phytohormones and phytometabolites in photosynthetic resilience, but here in this chapter we have compiled all these reports and discussed their roles. Besides, bioinformatic study of gamma TMT in cyanobacteria has also been done.
Preprint
Full-text available
Clonal plants, like cyanobacteria, are widespread and perform important ecosystem functions, influencing the structure and composition of the habitats in which they occur. Some cyanobacteria perform biological nitrogen fixation (BNF) and can affect plant growth as nitrogen (N) is a limiting nutrient. Therefore, to investigate whether heterocystous cyanobacteria favour individual growth and reproductive strategies (sexual reproduction and clonal growth) of Salvinia auriculata , we carried out a greenhouse experiment with the inoculation of two strains of cyanobacteria, Desmonostoc (UFLA 12) and Cronbergia (UFLA 35). S. auriculata ramets were grown in plastic pots with the following treatments: (D) Desmonostoc (UFLA 12) inoculum; (C) Cronbergia (UFLA 35) inoculum; (D + C) Cronbergia (UFLA 35) + Desmonostoc (UFLA 12) inoculum, and (Co) control, absence of cyanobacteria. Treatments (D) and (D + C) positively influenced the clonal growth of S. auriculata . Desmonostoc inoculation contributed to numerical increase in shoots, biomass gain, and shoot size. Cronbergia (UFLA 35) alone was not able to promote the growth of S. auriculata , only in consortium with Desmonostoc (UFLA 12). We conclude that the inoculation of Desmonostoc (UFLA 12) and Cronbergia (UFLA 35) favours the clonal growth of S. auriculata contributing to its more vigorous spread. The fact that Desmonostoc (UFLA 12) and Cronbergia (UFLA 35) favoured the clonal growth of S. auriculata may serve as a tool to assist in understanding the excessive growth of these plants in aquatic environments, for acting as a potential biofertiliser.
Article
Full-text available
Synthesizing nanoparticles through a green synthesis approach is common nowadays. Cyanobacteria have attained great importance in the field of biosynthesis of nanoparticles as there is no use of toxic chemicals as reducing or capping agents for the synthesis of metal oxide nanoparticles. Micronutrient-based nano-formulations have become a topic of great interest in recent times due to their various advantageous properties and applications in agriculture. The current study aims to exploit the potential cyanobacterial strains isolated from different locations such as freshwater and soil ecosystems. The potential cyanobacterial isolates were screened based on their multiple plant growth promoting (PGP) attributes such as Indol acetic acid (IAA) production, siderophores, and phosphate solubilization. After the screening of cyanobacteria based on multiple PGP activities, the cyanobacterial strain was identified at the species level as Pseudanabaena foetida RJ1, based on microscopy and molecular characterization using 16S rRNA gene sequencing. The cyanobacterial biomass extract and cell-free extracts are utilized for the synthesis of CuO micronutrient Nanoparticles (NPs). The cyanobacterial strain Pseudanabaena foetida RJ1 possesses plant growth-promoting (PGP) attributes that provide reduction and capping for CuO NPs. The synthesized NPs were characterized and subjected to make a nano-formulation, utilizing the cyanobacteria-mediated CuO NPs along with low-cost zeolite as an adsorbent. The application of cyanobacterial biomass extract and cell-free extract provided an excellent comparative aspect in terms of micronutrient NP synthesis. The NPs in the form of formulations were applied to germinated paddy seeds (Pusa Basmati -1509) with varying concentrations (5, 10, 15 mg/l). Effects of cyanobacteria based CuO NPs on hydroponically grown paddy crops were analyzed. The application of nano-formulations has shown a significant increase in plant growth promotion in rice plants under hydroponics conditions. There is no such type of comparative investigation reported earlier, and NPs of micronutrients can be utilized as a new economic nanofertilizer and can be applied to plants for their growth promotion.
Chapter
The chapter provides a comprehensive examination of the concept of secondary metabolites, encompassing discussions on various classification systems for these compounds, and an assessment of the available data management tools related to secondary metabolites. The main and most studied toxic and nontoxic secondary metabolites produced by cyanobacteria are described. Toxic metabolites are interpreted from the point of view of the primary target on which they act, while nontoxic metabolites are interpreted from the point of view of the biological role they play in cyanobacteria cells. For both categories, the possibility of practical application is being considered. Much attention is paid to cyanobacterial cytotoxins, which are promising substances in terms of creating drugs for the treatment of various forms of cancer. The chapter emphasizes the value of cyanobacteria as sources of valuable products, as well as their toxic potential in relation to other living organisms.
Article
summaryGunnera L. develops a complex and intimate symbiosis with N2-fixing cyanobacteria of the genus Nostoc, which renders the plant independent of combined nitrogen. The Nostoc-Gunnera symbiosis exhibits unique features compared to other cyanobacterial-plant symbioses: it is for example the only one that involves a flowering plant (angiosperm), the cyanobacterium infects specialized gland organs located on the stems of the host and once it has passed into the interior of the gland the cyanobacterium also enters the Gunnera cells where it starts to differentiate the highest frequency of heterocysts (the N2-fixing cells) recorded in any cyanobacterial population. Gunnera has attracted scientific attention also for the following reasons: the genus has a peculiar geographic distribution of its subgenera and species in the Southern Hemisphere. It differs morphologically and anatomically from related plants and also shows an anomalous polystelic vascular system (polystely). This review gives an updated account of present knowledge concerning the Nostoc-Gunnera symbiosis. Emphasis will be on the development of the symbiotic tissue (the gland), the structure and function of the prokaryotic N2-fixing cyanobacterium, the infection process and on the relationship between the pro- and eukaryotic partners prior to and following the establishment of symbiosis.
Article
A specific solid-phase enzyme immunoassay for the detection of as little as 3-4 pg of indole-3-acetic acid (IAA) is described. The assay involves minimal procedural efforts and requires only standard laboratory equipment. Up to 50 samples in triplicate, processed simultaneously, can be assayed and evaluated in 2.5 h. As little as 1 mg oat coleoptile tissue is sufficient for a quantitative IAA analysis and little or no extract purification is necessary. Using this assay, levels of IAA have been determined in coleoptiles of maize and oat. The distribution of IAA within single coleoptiles was quantitated and the production of IAA during the regeneration of the physiological tip in Avena coleoptiles was investigated. The changes in levels of IAA and other major phytohormones were quantitated during the growth of oat coleoptiles.
Article
Since the mid-1960s, following a century of unrewarded search, impressive progress has been made toward deciphering the Precambrian fossil record, evidence of life extant during the earliest seven-eighths of geologic time. Hundreds of fossiliferous units have been discovered containing thousands of microbial fossils—dominantly but not exclusively cyanobacterial — and the documented antiquity of life has been extended to an age roughly three-quarters that of the Earth. Mutually reinforcing lines of evidence, paleontological, geological, and isotopic geochemical, indicate that stromatoliticmicrobial ecosystems,evidently including cyanobacteria and other members of the bacterial domain, were extant ~3500 Ma ago; methanogenic archaeans by ~2800 Ma ago; and Gram-negative sulfate-reducing bacteria at least as early as ~2700 Ma ago.The discrepancy between these dates and those suggested for emergence of these groups by a recently proposed amino acid-based “molecular clock” is too great and too consistent to be ignored. The challengeis to unify the molecular data with the increasingly well-established paleobiologicrecord.
Article
The Anthoceros punctatus-Nostoc and Blasia pusilla-Nostoc symbioses were investigated. In both associations the Nostoc colonies develop in mucilaginous cavities on the undersurface of the gametophyte. The sporophytes of Anthoceros have no Nostoc colonies; no sporophytes of Blasia were found. The symbiotic algae have been identified as Nostoc sphaericum ex. Born et Flah. The heterocyst frequency of the free-living isolates is 3–6% but this increases to 30% (Blasia) and 43% (Anthoceros) when the algae are growing symbiotically. On infecting alga-free gametophytes of Blasia pusilla with Nostoc, algal colonies develop in the cavities within 72 h. The developing Nostoc colonies stretch the cells of the cavity wall and there also arise from a point on the host cavity wall, usually opposite the cavity pore, septate, branched, filamentous protrusions, which increase the surface area of contact between phycobiont and host by about 30% within 4 weeks of colony formation. Such outgrowths may facilitate interchange of metabolites between the alga and liverwort. The Nostoc phycobiont of Anthoceros may infect Blasia and vice-versa. A strain of Nostoc punctiforme isolated from Gunnera also infects Blasia, but four other Nostoc isolates, including one from cycad root nodules, do not. Five other heterocystous algae, five non-heterocystous filamentous algae and one unicellular alga tested do not infect Blasia. The nitrogen contents of thalli with Nostoc colonies are significantly higher than those treated with algae which do not develop colonies. The symbiotic algae differ physiologically and biochemically from free-living algae. They show higher rates of nitrogenase activity (acetylene reduction assay), are depleted of nitrogen-storing phycobilin pigments and structured granules and although metabolically active do not evolve O2 or fix CO2 photosynthetically, but do have polyhedral bodies which contain the key CO2-fixing enzyme ribulose-l,5-diphosphate carboxylase. The Blasia symbiosis grows well and remains established in the pH range 4–8, at low levels of combined nitrogen (
Article
The enigmatic coexistence of O2-sensitive nitrogenase and O2-evolving photosynthesis in diazotrophic cyanobacteria has fascinated researchers for over two decades. Research efforts in the past 10 years have revealed a range of O2 sensitivity of nitrogenase in different strains of cyanobacteria and a variety of adaptations for the protection of nitrogenase from damage by both atmospheric and photosynthetic sources of O2. The most complex and apparently most efficient mechanisms for the protection of nitrogenase are incorporated in the heterocysts, the N2-fixing cells of cyanobacteria. Genetic studies indicate that the controls of heterocyst development and nitrogenase synthesis are closely interrelated and that the expression of N2 fixation (nif) genes is regulated by pO2.
Article
The production and characterization of high-affinity monoclonal antibodies suitable for the radio- and enzymeimmunoassay of the endogenous plant growth regulator, indole-3-acetic acid (IAA), is reported. Hybridomas were produced by fusion of NS 1 myeloma cells with spleen cells from Balb/c mice immunized with IAA-bovine serum albumin conjugates. From an initial collection of 158 wells containing cells secreting monoclonal antibodies against IAA, seven were used to derive cell clones. Three of these are described here. They secrete immunoglobulin (IgG2a or IgG2b) of high affinity and specificity for IAA methyl ester and can be used to quantite picogram amounts of this compound in plant extracts by radio- and enzymeimmunoassay.