You are on page 1of 10

520 Protein & Peptide Letters, 2012, 19, 520-529

Antimicrobial Peptides from Adenanthera pavonina L. Seeds: Characteri-


zation and Antifungal Activity

Julia Ribeiro Soares1,¶, André de Oliveira Carvalho1,¶, Izabela Silva dos Santos1, Olga Lima Tavares
Machado2, Viviane V. Nascimento1, Ilka Maria Vasconcelos3, André Teixeira da Silva Ferreira4,
Jonas Enrique de Aguilar Perales4 and Valdirene Moreira Gomes1,*

1
Laboratório de Fisiologia e Bioquímica de Microrganismos, Universidade Estadual do Norte Fluminense, Campos dos
Goytacazes, RJ, Brazil; 2Laboratório de Química e Função de Proteínas e Peptídeos, Universidade Estadual do Norte
Fluminense, Campos dos Goytacazes, RJ, Brazil; 3Laboratório de Toxinas Vegetais, Universidade Federal do Ceará,
CE, Brazil; 4Oswaldo Cruz, Rio de Janeiro, RJ, Brazil

Abstract: In this study, the antifungal activity of peptides extracted from Adenanthera pavonina seeds was assessed. Pep-
tides were extracted and fractionated by DEAE-Sepharose chromatography. The non-retained D1 fraction efficiently in-
hibited the growth of the pathogenic fungi. This fraction was later further fractionated by reversed-phase chromatography,
resulting in 23 sub-fractions. All separation processes were monitored by tricine-SDS-PAGE. Fractions H11 and H22
strongly inhibited the growth of Saccharomyces cerevisiae and Candida albicans. Fraction H11 caused 100% death in S.
cerevisiae in an antimicrobial assay. The complete amino acid sequence of the peptide in fraction P2 was determined, re-
vealing homology to plant defensins, which was named ApDef1. Peptides from fraction H22 were also sequenced.
Keywords: Adenanthera pavonina, Candida albicans, defensin, fungicide activity, membrane permeabilization, Saccharomy-
ces cerevisiae.

INTRODUCTION [13], snakins [14] and cyclotides [15]. These antimicrobial


substances are particularly abundant in seeds [16], where
Innate immunity is important for the survival of multicel-
they are constitutively expressed during normal development
lular organisms under threat by invading microorganisms. A and participate in protection against microorganisms in soil
key component of innate host defense is the production of
[17].
antimicrobial peptides (AMPs) [1]. Indeed, virtually every
form of life, from bacteria on, produces these self-defense In recent decades, AMPs have been considered a promis-
molecules [1]. AMPs are low molecular weight amphiphilic ing source of new pharmaceuticals [18], particularly given
peptides that contain a positive net charge at physiological increased bacterial resistance to conventional antibiotics
pH and are endowed with antimicrobial activity [2]. AMPs [19], inadequate investment in the development of new anti-
may be constitutively expressed or inducibly produced upon biotics [20] and increased cases of invasive mycoses in im-
sensing an invader, and their antimicrobial effect is primarily munocompromised patients, primarily those caused by Can-
mediated through interaction with biological membranes. In dida spp. [21]. In particular, plant defensins have been sug-
addition to this primary target, intracellular targets have al- gested as a possible new source of therapeutics to treat inva-
ready been described for some AMPs, including binding to sive mycoses [22].
chaperones [3], interfering with macromolecular synthesis Previous studies demonstrated that A. pavonina seeds
[4] and disrupting the cell cycle by interacting with cyclin F contain a chitinase with robust thermal stability [23, 24]. The
[5]. For the majority of AMPs, the mechanism of action present report contains an analysis of the AMPs present in A.
leading to microbial growth inhibition remains unknown. pavonina seeds. The purification of a peptide with homology
In plants, several families of AMPs have been described to the plant defensin family that exhibits potent inhibitory
and characterized; most of these families are rich in disul- activity against S. cerevisiae is reported.
fide-forming Cys residues. These disulfide bonds endow
AMPs with great physicochemical stability [6,7]. Plant- MATERIAL AND METHODS
derived AMPs include lipid transfer proteins [8, 9], thionins
Plant Materials
[10], plant defensins [11], hevein-like peptides, knottin-like
peptides, glycine-rich peptides [12], MBP-1 homologs Seeds from Adenanthera pavonina (L.) were collected in
fields of Campos dos Goytacazes province, Rio de Janeiro,
*Address correspondence to this author at the Laboratório de Fisiologia e
Brazil.
Bioquímica de Microrganismos, Centro de Biociências e Biotecnologia,
Universidade Estadual do Norte Fluminense, Campos dos Goytacazes, Fungi
28013-602, RJ, Brazil; Tel: +55 22 2739-7033; Fax: +552227397178;
E-mail: valmg@uenf.br Saccharomyces cerevisiae (1038), Candida albicans

These authors contributed equally to this work. (CE022) and Candida tropicalis (CE017) were obtained

-5/12 $58.00+.00 © 2012 Bentham Science Publishers


Antimicrobial Peptides from Adenanthera pavonina Protein & Peptide Letters, 2012, Vol. 19, No. 5 521

from the Departamento de Biologia, Universidade Federal do Amino Acid Sequence Analysis
Ceará, Fortaleza, Brazil. Fusarium oxysporum (FBM 002)
Amino acid sequence analysis was performed on peptides
was kindly provided by CNPAF/EMBRAPA, Goiania,
isolated from fractions H11 and H22. These fractions were
Goiás, Brazil. Fungi were maintained on Sabouraud agar
treated with 5 mM -mercaptoethanol (Sigma) and 5%
(1% peptone, 2% glucose and 1.7% agar-agar).
vinylpyridine (Sigma) and were subsequently heated for 30
min at 37 ºC and 15 min at 60 ºC. After treatment, the sam-
Extraction and Purification of Antimicrobial Peptides ples were separated by tricine-SDS-PAGE, transferred to
Purification of peptides from A. pavonina seeds was per- polyvinylidene difluoride (PVDF, Millipore) membranes and
formed as described previously with modifications [25]. Pep- stained with Ponceau-S (0.1%, GE HealthCare) [28]. The
tides from seed flour (50 g) were extracted for 3 h (4 ºC) band corresponding to approximately 7 kDa from fraction
with 500 mL of extraction buffer (10 mM Na2HPO4, 15 mM H11 and the bands at approximately 3 and 6 kDa from frac-
NaH2PO4, 100 mM KCl and 1.5% EDTA, pH 5.4). The pre- tion H22 were excised from the corresponding membrane;
cipitate formed between 0 and 90% relative ammonium sul- each was briefly washed three times in water, and the mem-
fate saturation, after centrifugation (15000 x g, 30 min), was brane was air-dried. N-terminal amino acid sequences of
redissolved in distilled water and heated at 80 ºC for 15 min. peptides blotted onto PVDF were determined by Edman deg-
The resulting suspension was clarified by centrifugation radation carried out in a Shimadzu PSQ-23A protein se-
(10000 x g, 10 min) and the supernatant was extensively quencer (Shimadzu). Phenylthiohydantoin (PTH) –derivati-
dialyzed against distilled water. The dialyzed solution was zed amino acids were detected at 269 nm after separation on
recovered by freeze drying (fraction F/0-90) and was further a reversed-phase C18 column (250 x 4.6 mm) under isocratic
used for antimicrobial peptides purification. conditions according to the manufacturer's instructions.
Searches for sequence homology were performed using
Separation of proteins from fraction F/0-90 was carried
BLASTp [29], and selected sequences were aligned with
out using a DEAE-Sepharose column (3 x 7 cm, GE
ClustalW [30].
HealthCare) equilibrated with 100 mM Tris-HCl (pH 8.0).
The column was eluted at a flow rate of 90 mL/h, first with The complete sequence of the lower molecular weight
equilibrium buffer to obtain fraction D1 and then with equi- peptide of fraction H11, referred to as P2, was determined by
librium buffer containing 1 M NaCl to obtain bound proteins proteolytic digestion followed by N-terminal sequencing.
(fraction D2). Protein elution was monitored by measurement The purified peptide was dissolved in 100 mM Tris-HCl, pH
of absorbance at 280 nm (Spectrophotometer LGS53, BEL 7.8, the endoproteinase Glu-C (Sigma) was added at a 1:50
Photonics). ratio (μg:μg) (H11 endoproteinase:sample), and the solution
was incubated at 37 ºC for 16 h. After this treatment, the Cys
Fraction D1 was recovered by freeze drying (FreeZone
residues in the peptide fragments were reduced and purified
4.5, Labconco), redissolved in 0.1% (v/v) trifluoroacetic acid
by reversed-phase HPLC as described previously using a
(TFA, Fluka) and loaded onto an HPLC (Prominence, Shi-
C18 column attached to a C8 guard column. The solvent
madzu) C18 reversed-phase column (250 x 4.6 mm, Shima-
flow rate was 0.5 mL/min, and the following solvent gradient
dzu) attached to a C8 pre-column (20 x 4.6 mm, Pelliguard, was used: 100% solvent A for 10 min, 0 to 60% solvent B
Sigma-Aldrich). The solvent flow rate was 0.5 mL/min, and
for 130 min, 60 to 100% solvent B for 5 min and 0% solvent
the following solvent gradient was used: 100% solvent A
B for 15 min. Elution was monitored by online measurement
(0.1% TFA in water) for 7 min, 0 to 42% solvent B (100% 2-
of absorbance at 220 and 280 nm. The eluted sample was
propanol containing 0.1% TFA) for 64 min, 50% solvent B
dried in a Speed Vac Plus (SC110A, Savant). Two main
for 2 min and 0% solvent B for 10 min. Elution was moni-
peaks were obtained after cleavage. The corresponding frac-
tored by online measurement of absorbance at 220 and 280 tions were subjected to N-terminal amino acid sequence
nm.
analysis, and sequence homology searches were performed
Fraction H11 was diluted in 0.1% (v/v) trifluoroacetic as described above. In addition, the purified peptide, in both
acid and loaded onto an HPLC C2C18 reversed-phase col- native form and reduced, alkylated form, was subjected to
umn (100 x 4.6 mm, μRPC ST 4.6/100, GE HealthCare). The mass spectrometry analysis. Peptide was cocrystalized with
solvent flow rate was 0.5 mL/min, and the following solvent an excess of the matrix alpha cyano-4-hydroxycinnamic acid
gradient was used: 100% solvent A for 5 min, 0 to 17% sol- and further analyzed by matrix-assisted laser desorption ioni-
vent B for 10 min, 22% solvent B for 10 min and 0% solvent sation time-of-flight mass spectrometry (MALDI-TOF MS).
B for 10 min. Elution was monitored by online measurement The instrument used was an AB SCIEX TOF/TOF™ 5800
of absorbance at 220 and 280 nm. System spectrometer (AB SCIEX) in the linear mode.
Protein content was determined as described by Bradford
[26] with albumin from chicken egg white (Sigma) used as Effect of Fraction D1, H11 and H22 on Fungi Growth
the protein standard. Fungus and yeast cultures were prepared by transferring
an inoculum of each to Petri dishes containing Sabouraud
Electrophoresis agar and allowing them to grow at 30 ºC for 48 h. After this
period, the cells were transferred to sterile Sabouraud broth
Tricine-sodium dodecyl sulfate-polyacrylamide gel elec-
(1 mL). Yeast cells were counted in a Neubauer chamber for
trophoresis (tricine-SDS-PAGE) was performed according to
further calculation of appropriate dilutions. To monitor the
the method of Schägger and von Jagow [27].
effect of AMPs from A. pavonia seeds on the growth of
yeasts, 104 cells/mL (in Sabouraud broth) were incubated in
522 Protein & Peptide Letters, 2012, Vol. 19, No. 5 Soares et al.

the presence of 25, 50 and 100 μg/mL peptide at 28 ºC in 450-490 nm; emission 500 nm). An untreated negative con-
200 μL total volume in 96 well microplates. Culture medium trol was included to evaluate the baseline membrane perme-
was used as the blank for the spectrophotometer. Optical ability, and a positive control was included that contained
readings were taken at 670 nm at the zero time point and cells killed by heating at 100 ºC for 5 min.
every 6 h for the following 60 h for fungi and 36 h for yeasts
[31]. Optical densities were plotted as a function of absor- RESULTS
bances.
Extraction and Partial Purification of A. pavonina An-
timicrobial Peptides
Effect of Fractions H11 and H22 on Yeast Cell Viability
Fraction F/0-90 from A. pavovina seeds was separated
After yeast cultures were grown as described above for
into two different fractions, D1 and D2, by anion exchange
36 h, cells in the untreated control wells were quantified in a
chromatography on a DEAE-Sepharose column Fig. (1).
Neubauer chamber as a control dilution. The untreated con-
trol and test samples (treated with 80 μg/mL of each frac- Tricine-SDS-PAGE analysis showed that these fractions
contained peptides with molecular weights between 6 and 10
tion) were diluted in Sabouraud broth as appropriate and
kDa and that they contained several proteins with molecular
plated with a Drigalski loop on Petri dishes containing
weights above 16 kDa (Fig. 1 insert).
Sabouraud agar. Cells were evenly distributed on the plate
and were incubated at 30 ºC for 48 h to allow colony devel- Fraction D1 was further fractionated into 23 fractions by
opment. After this period, the colonies were counted. Data reversed-phase chromatography (RP) on a C18 column; one
were analyzed with respect to the untreated control sample, fraction contained the flow-through, and the other 22 frac-
which was considered to be 100% viable. tions contained material eluted from a propanol gradient Fig.
(2). With the exception of some peptides in fraction 22, all
Plasma Membrane Permeabilization peptides detected were between 2 and 17 kDa Fig. (3).
Among the 22 isopropanol gradient fractions, two criteria
Plasma membrane permeabilization was measured by were used to choose the fractions that were tested for antimi-
SYTOX® Green uptake as described previously by Thevis- crobial activity: the homogeneity of the fraction and the
sen et al. [32] with modifications. After fungal and yeast presence of low molecular weight peptides. Using these cri-
cultures were grown for 36 h as described above, 100 μL teria, fractions H11, H17, H18, H19, H21 and H22 were selected
aliquots of S. cerevisiae, C. albicans and C. tropicalis sus- for further analysis. Only fractions H11 and H22 exhibited
pensions that were grown in the presence of the AMP- antimicrobial activity (results not shown for other fractions).
containing fractions from A. pavonina were further incubated
with 0.2 μM SYTOX® Green in 96 well microplates for 1.5
Characterization of the Antimicrobial Activity
h at 25 ºC with periodic agitation. Cells were then observed
using a DIC microscope (Axiophoto Zeiss) equipped with a The antimicrobial assay demonstrated that the fraction D1
fluorescence filter set for fluorescein detection (excitation possessed antimicrobial activity, which was expected

Figure 1. Chromatogram of the fraction F/0-90 in DEAE-Sepharose column. D1 was eluted with equilibrium buffer (Tris-HCl 0.1 M, pH
8.0), and D2 was eluted with equilibrium buffer containing 1 M NaCl. The arrow indicates the buffer change. Insert: Electrophoretic visuali-
zation of fraction F/0-90 and fractions D1 and D2 in tricine-SDS-PAGE. F/0-90, fraction obtained after 90% relative ammonium sulfate satu-
ration. Fraction D1, not retained in the DEAE-Sepharose column; fraction D2, retained in the DEAE-Sepharose column and eluted with 1 M
NaCl; M, molecular weight markers indicated in kDa.
Antimicrobial Peptides from Adenanthera pavonina Protein & Peptide Letters, 2012, Vol. 19, No. 5 523

Figure 2. C18 reversed-phase HPLC chromatogram of fraction D1. D1 was fractionated into 23 fractions using a propanol/TFA gradient
(oblique line) at a flow rate of 0.5 mL/min. Fractions were named as unbound and H1 to H22. Elution was monitored at 220 nm.

because of the basic character of the peptides present in this


fraction. At the endpoint of the antimicrobial assay, fraction
D1 demonstrated approximately 37, 53 and 82% growth in-
hibition of the fungus F. oxysporum at 25, 50 and 100
μg/mL, respectively Fig. (4A). Fraction D1 did not inhibit
growth of S. cerevisiae Fig. (4B). At the endpoint of the as-
say, growth inhibition of C. tropicalis was 32 and 79% at the
concentrations 50 and 100 μg/mL, respectively Fig. (4C).
Inhibition of C. tropicalis growth was not observed at 25
μg/mL.
At 80 μg/mL, fraction H11 inhibited growth of the yeasts
S. cerevisiae and C. albicans by approximately 83% and
42%, respectively at the assay endpoint Fig. (5A and 5B).
Fraction H22 inhibited growth of S. cerevisiae by approxi-
mately 22% at a concentration of 80 μg/mL Fig. (6A) and
did not inhibit growth of C. albicans Fig. (6B).
To characterize the mechanism of the antimicrobial ef-
fects of the H11 and H22 fractions, we also tested the effects
of these fractions on yeast membrane permeabilization and
cell viability. The membrane permeabilization assay was
performed immediately after the growth assay for the yeasts
S. cerevisiae and C. albicans, using the nucleic acid stain
SYTOX® Green. Cells treated with fractions H11 and H22 did
not exhibit an increase in fluorescence when compared with
the untreated control, indicating that the membranes of these
cells were not permeabilized by the peptides present in the
fractions (Table 1).
After the growth assay, cells from the control were
counted, diluted and plated on Sabouraud agar. Similarly,
cells treated with the H11 fraction were diluted and plated. S.
cerevisiae cells treated with fraction H11 exhibited 100%
death Fig. (7A and B), whereas C. albicans cells exhibited
0% death Fig. (7C and D).
Purification of Peptides of H11 and Amino Acid Sequence
Figure 3. Electrophoretic visualization of the 22 bound fractions in
of P2 and H22 Fractions
reversed-phase C18 column in tricine-SDS-PAGE. 1 to 22, bound To identify the peptides contained in fraction H11, the
fractions and eluted upon propanol gradient formation. M, molecu- mixture was submitted to C2C18 RP chromatography, result-
lar weight markers are indicated in kDa.
524 Protein & Peptide Letters, 2012, Vol. 19, No. 5 Soares et al.

Figure 5. Growth curves of the yeasts Saccharomyces cerevisiae


(A) and Candida albicans (B) in the presence of fraction H11. ()
control; () 80 g/mL, (n = 3).

Figure 4. Growth curves of the filamentous fungus Fusarium ox-


ysporum (A) and the yeasts Saccharomyces cerevisiae (B) and
Candida tropicalis (C) in the presence of fraction D1. () control;
() 25 g/mL; () 50 g/mL and () 100 g/mL, (n = 3).

ing in three peaks Fig. (8). Peak P2 contained a single peptide


of approximately 7 kDa, which corresponds to the ApDef1 ,
and P3 contained a single peptide of approximately 17 kDa
(Fig. (8) insert).
N-terminal amino acid sequence analysis of the lower
molecular weight peptide of fraction H11, approximately 7
kDa in size, showed a homology with the primary structure
of antimicrobial peptides from the plant defensin family (re-
sult not shown). To obtain the complete amino acid sequence
of P2, the purified sample was digested with endo Glu-C, and
the proteolytic fragments were purified. Two main peaks
were obtained after cleavage. These two peaks were sub-
jected to N-terminal amino acid sequence analysis, and pro-
tein sequence databases were queried to identify similar pep-
tides. This analysis revealed that we had determined the first
31 and the last 17 amino acids of a peptide from plant de-
fensin family. The first 31 amino acids were determined to Figure 6. Growth curves of the yeasts Saccharomyces cerevisiae
arise directly from the peptidyl N-terminus. By combining (A) and Candida albicans (B) in the presence of the fraction H22.
() control; () 80 g/mL, (n = 3).
Antimicrobial Peptides from Adenanthera pavonina Protein & Peptide Letters, 2012, Vol. 19, No. 5 525

analyses to characterize this peptide more fully are under-


way.

DISCUSSION
Most AMPs are characterized by low molecular weight,
amphipathicity and basic character. A purification strategy
was developed to isolate peptides with these properties from
A. pavonina seeds based on an extraction and purification
method that was previously used to isolate AMPs from seeds
of Raphanus sativus [25]. Other researchers have used simi-
lar approaches, such as a combination ion exchange and re-
versed-phase (RP) chromatography separations, to purify
AMPs from protein extracts of plant tissue or organs [16, 25,
38]. Other combinations of chromatographic techniques have
also been utilized, including ion exchange and gel filtration
[39, 40] or RP alone [6, 41]. In each case, the purification
method used depended on the charge, size, physical stability,
and other properties of the desired peptides. We obtained 22
bound fractions using this approach; of the 6 fractions se-
lected for further analysis, 2 contained antimicrobial activity.
Electrophoretic analysis showed that fraction H11 contained
one major peptide of approximately 7 kDa and that fraction
H22 contained peptides ranging from 2 to 10 kDa Fig. (3).
Figure 7. Cell viability assay of Saccharomyces cerevisiae (A, Fraction H11 was subjected to C2C18 RP chromatogra-
control and B, test) and Candida albicans (C, control and D, test) phy, resulting in three peaks Fig. (8). The purified peptide in
treated with fraction H11. peak P2 was sequenced, revealing a homology to the primary
structure of the plant defensin family of antimicrobial pep-
this information with the database query results and the tides. Plant defensins are short, basic peptides composed of
known substrate specificity of endo Glu-C, we were able to 45 to 54 amino acids that feature a structure of three anti-
determine the endo Glu-C cleavage site and the complete parallel -strands and one -helix [42]. The structure is sta-
peptide sequence. In addition, we obtained a mass of 5129 bilized by four disulfide bonds that form a cysteine-
Da for the intact peptide and 5590 Da for the intact alkylated stabilized  motif characteristic of these proteins [43, 44].
peptide, consistent with the presence of eight Cys residues The presence of eight Cys residues is strictly conserved in
(result not shown). The amino acid sequence of P2 exhibited primary structures of plant defensins. Among the 48 amino
a 72, 41 and 42% homology with the defensins 5144 (from acid residues sequenced, Cys residues were identified at po-
Cassia fistula), PsD2 and NsD1, respectively Fig. (9A). sitions 3, 14, 23, 27, 38, 42, 44 and 48. The presence of eight
The 21 amino acid N-terminal sequence obtained from Cys residues was further confirmed by mass spectrometry
the 6 kDa peptide contained in fraction H22 showed a homol- analysis. The only known exception to the requirement for
ogy with the primary structure of the ubiquitin family. It eight Cys residues in members of the plant defensin family is
exhibited 95% identity with ubiquitin from Oryza sativa, the peptide Ph1, which has 10 Cys residues and five disulfide
Arabidopsis thaliana and Daucus carota Fig. (9B). The 18 bridges [45]. The comparison also indicated that among the
amino acid N-terminal sequence obtained from the 3 kDa plant defensin family, other amino acid residues were also
peptide (SNYDYDVETETAYAYT) contained in fraction conserved, such as an aromatic residue at position 10, Gly12,
H22 revealed no sequence homology with any other known Glu31 and Gly36 Fig. (9A). In the future, the peptide identi-
peptide, even those isolated from other plant seeds. Further
526 Protein & Peptide Letters, 2012, Vol. 19, No. 5 Soares et al.

Figure 8. C2C18 reversed-phase HPLC chromatogram of fraction H11. H11 was fractionated into 4 fractions using a propanol/TFA gradient
(oblique line) and at a flow rate of 0.5 mL/min. Fractions were named as P1 to P4. Elution was monitored at 220 nm. Insert: Electrophoretic
visualization of the three bound fractions separated by C2C18 reversed-phase HPLC in tricine-SDS-PAGE. P1, P2 and P3, bound fractions
and eluted with a propanol gradient. M, molecular weight markers are indicated in kDa.

fied in this work will be referred to as ApDef1 (A. pavonina microbial membranes, including plant defensins [32, 53],
defensin one). temporins [54], plant lipid transfer protein [55] and the knot-
tin-type peptide psacotheasin [56]. Interestingly, ApDef1 did
Like insect and mammal defensins, plant defensins pos-
not cause membrane permeabilization of the yeasts tested in
sess antimicrobial activity. In plant defensins, this activity is
this study (Table 1), but the peptide killed 100% of the S.
directed primarily against fungal pathogens [38, 46]. The
cerevisiae cells in a viability assay Fig. (7). The results indi-
following two yeast strains were selected for antimicrobial
activity testing in this study: S. cerevisiae, which is a model cated that fraction H11 exhibited a fungicidal effect on S.
cerevisiae and a fungistatic effect on C. albicans. In addi-
yeast featuring the availability of myriad genetic tools, ex-
tion, a mechanism of action other than membrane permeabi-
tensive knowledge of its cell physiology and broad annota-
lization may be involved because ApDef1 did not cause gross
tion of its genome [47]; and C. albicans, which is an oppor-
plasma membrane disorganization, at least for S. cerevisiae.
tunistic human pathogen causing life-threatening systemic
There are several known examples of AMPs for which mem-
diseases in immunocompromised patients in whom control
of the pathogen has been difficult [48]. The fraction contain- brane permeabilization was not the mechanism of action
responsible for microbial growth inhibition or death. These
ing ApDef1 inhibited 83% of S. cerevisiae growth and 42%
examples include human  defenisn-2 (C. albicans) [57],
of C. albicans growth Fig. (5).
histatin-5 and (C. albicans) [58] and dermicidin-derived pep-
Some plant defensins have been shown to preferentially tides (bacteria) [59]. Aerts et al. [60] demonstrated that the
bind to sphingolipids in fungal membranes; Rs-AFP2 and defensin from Raphanus sativus, Rs-AFP2, did not permeabi-
MsDEF1, are known to bind glucosylceramides (GlcCer), lize small unilamellar vesicles composed of GlcCer in the
whereas Dm-AMP1 binds mannosyldiinositol phosphorylce- carboxyfluorescein leakage assay. The authors also demon-
ramide (M(IP)2C) [49, 50]. Fungal membranes contain sev- strated the production of reactive oxygen species (ROS) and
eral different classes of sphingolipids. For example, S. cere- concluded that the inhibition effect was derived from an in-
visiae membranes contain galactosylceramide and M(IP)2C tracellular mechanism.
[51, 52] but not GlcCer, whereas C. albicans membranes do
contain GlcCer [52]. The preferential binding of plant de- It has been demonstrated that some plant defensins are
internalized into the fungal cytoplasm [5, 53]. For defensin
fensins to a particular sphingolipid class, together with the
PsD1, the mechanism of fungal growth arrest appeared to be
differential presence of the sphingolipid classes in different
related to the interaction of internalized defensin with intra-
yeast membranes, may help explain why ApDef1 killed S.
cellular targets. In this case, once inside the cell, PsD1 mi-
cerevisiae but only inhibited the growth of C. albicans Fig.
grated to the nucleus and interacted with cyclin F, interfering
(7).
with the fungal cell cycle [5]. Other mechanisms were also
The antimicrobial activity of AMPs is believed to be described, such as production of ROS [55, 61] and the induc-
caused by their interaction with microbial membranes. Sev- tion of apoptosis [60].
eral AMP classes have been demonstrated to permeabilize
Antimicrobial Peptides from Adenanthera pavonina Protein & Peptide Letters, 2012, Vol. 19, No. 5 527

Figure 9. Comparison of the complete amino acid sequence obtained from ApDef1 with plant defensins (A) and of the 21 amino acid N-
terminal sequence of the 6 kDa peptide from fraction H22 with ubiquitins (B). Sequences were aligned with the ClustalW [30]. I% indicates
the percentage of identical residues, which are shown in italics, and P% indicates the percentage of positive residues (representing amino
acids with the same biochemical characteristics such as charge and hydrophobicity), which are shown in gray. Gaps (-) were included to
optimize alignment. The numbers above the sequence indicate peptide length in amino acids excluding gaps, and sequences were numbered
according to ApDef1. The numbers flanking the sequences indicate the first N-terminal amino acid and the last C-terminal amino acid. The
residue E marked with an asterisk indicates the point where the two peptides were joined. Sequences from the following defensins were pre-
sented, referring to the name of the peptide or gene identification number: For (A) Cassia fistula [33], VrD2 [34], Tad1 [35], EcAMP1 [36],
Triticum aestivum (Q8L698), NsD1 [37], DmAMP1 [38], Rs-AFP1 [33], PsD2 (PSD2P81930). For (B) Oryza sativa (GI 302393766), Arabi-
dopsis thaliana (GI 302393741), Hordeum vulgare (GI 302595831), Daucus carota (GI302393742), Helianthus annuus (GI302393790) and
Schizosaccharomyces pombe (GI 302393762).

[2] Yount, N.Y.; Yeaman, M.R. Multidimensional signatures in antim-


CONCLUSION icrobial peptides. Proc. Natl. Acad. Sci. USA, 2004, 101(19), 7363-
7368.
A peptide was obtained from an A. pavonina seed extract [3] Kragol, G.; Lovas, S.; Varadi, G.; Condie, B.A.; Hoffmann, R.;
that exhibited fungicidal effects on S. cerevisiae and inhib- Otvos, L. Jr. The antibacterial peptide pyrrhocoricin inhibits the
ited the growth of C. albicans. This peptide showed similar- ATPase actions of DnaK and prevents chaperone-assisted protein
folding. Biochemistry, 2001, 40, 3016-3026.
ity to plant defensin peptides, both in primary sequence ho- [4] Ulvatne, H.; Samuelsen, O.; Haukland, H.H.; Krämer, M.; Vorland,
mology and in antimicrobial activity. This newly discovered L.H. Lactoferricin B inhibits bacterial macromolecular synthesis in
peptide is named ApDef1. Future studies will examine the Escherichia coli and Bacillus subtilis. FEMS Microbiol. Lett.,
mechanism of action responsible for antifungal activity. In 2004, 237, 377-384.
addition, two other peptides from H22 fraction were se- [5] Lobo, D.S.; Pereira, I.B.; Fragel-Madeira, L.; Medeiros, L.N.;
Cabral, L.M.; Faria, J.; Bellio, M.; Campos, R.C.; Linden, R.; Kur-
quenced. A 6 kDa peptide showed sequence homology with tenbach, E. Antifungal Pisum sativum defensin 1 interacts with
ubiquitin, and a 3 kDa peptide did not exhibit homology with Neurospora crassa cyclin F related to the cell cycle. Biochemistry,
any other known peptide. The importance of this discovery is 2007, 46, 987-996.
under investigation. [6] Colgrave, M.L.; Craik, D.J. Thermal, chemical, and enzymatic
stability of the cyclotide kalata B1: the importance of the cyclic
cysteine knot. Biochemistry, 2004, 43, 5965-5975.
ACKNOWLEDGMENTS [7] Benko-Iseppon, A.M.; Galdino, S.L.; Calsa Jr, T.; Kido, E.A.;
Tossi, A.; Belarmino, L.C.; Crovella, S. Overview on plant antimi-
This study comprises part of the Graduation and M.Sc. crobial peptides. Curr. Protein Pept. Sci., 2010, 11, 181-188.
degree dissertation of JRS, first author, and was carried out [8] Carvalho, A.O.; Gomes, V.M. Role of plant lipid transfer proteins
at the Universidade Estadual do Norte Fluminense. We ac- in plant cell physiology - a concise review. Peptides, 2007, 28,
knowledge financial support from the Brazilian agencies 1144-1153.
[9] Padovan, L.; Segat, L.; Tossi, A.; Calsa Jr., T.; Ederson, A.K.;
CNPq and FAPERJ. We are also thankful for financial sup- Brandao, L.; Guimarães, R.L.; Pandolfi, V.; Pestana-Calsa, M.C.;
port to Dr. Viviane V. Nascimento through a fellowship to Belarmino, L.C.; Benko-Iseppon, A.M.; Crovella, S. Characteriza-
CAPES/PNPD. tion of a new defensin from cowpea (Vigna unguiculata (L.)
Walp.). Protein Pept. Lett., 2010, 17, 297-304.
[10] Bohlmann, H.; Apel, K. Thionins. Annu. Rev. Plant Physiol. Plant
REFERENCES Mol. Biol., 1991, 42, 227-240.
[1] Reddy, K.V.R.; Yedery, R.D.; Aranha, C. Antimicrobial peptides: [11] Carvalho, A.O.; Gomes, V.M. Plant defensins - prospects for the
premises and promises. Int. J. Antimicrob. Agents, 2004, 24, 536- biological functions and biotechnological properties. Peptides,
547. 2009, 30, 1007-1020.
528 Protein & Peptide Letters, 2012, Vol. 19, No. 5 Soares et al.

[12] Broekaert, W.F.; Cammue, B.P.A.; De Bolle, M.F.C.; Thevissen, [34] Lin, K.F.; Lee, P.H.; Hsu, M.P.; Chen, C.S.; Lyu, P.C. Structure-
K.; De Samblanx, G.; Osborn, R.W. Antimicrobial peptides from based protein engineering for -amylase inhibition activity of plant
plants. Crit. Rev. Plant Sci., 1997, 16, 297-323. defensin. Protein, 2007, 68, 530-540.
[13] Duvick, J.P.; Rood, T.; Rao, A.G.; Marshakg, D.R. Purification and [35] Koike, M.; Okamoto, T.; Tsuda, S.; Imai, R. A novel plant de-
characterization of a novel antimicrobial peptide from maize (Zea fensin-like gene of winter wheat is specifically induced during cold
mays L.) kernels. J. Biol. Chem., 1992, 267(26), 18814-18820. acclimation. Biochem. Biophys. Res. Commun., 2002, 98(2), 46-53.
[14] Berrocal-Lobo, M.; Segura, A.; Moreno, M.; López, G.; García- [36] Odintsova, T.I.; Rogozhin, E.A.; Baranov, Y.; Musolyamov, A.K.;
Olmedo, F.; Molina, A. Snakin-2, an antimicrobial peptide from Yalpani, N.; Egorov, T.A.; Grishin, E.V. Seed defensins of barn-
potato whose gene is locally induced by wounding and responds to yard grass Echinochloa crusgalli (L.) Beauv. Biochimie, 2008, 90,
pathogen infection. Plant Physiol., 2002, 128, 951-961. 1667-1673.
[15] Daly, N.L.; Rosengren, K.J.; Craik, D.J. Discovery, structure and [37] Rogozhin, E.A.; Oshchepkova, Y.I.; Odintsova, T.I.; Khadeeva,
biological activities of cyclotides. Adv. Drug Deliver. Rev., 2009, V.K.; Veshkurova, O.N.; Egorov, T.A.; Grishin, E.V.; Salikhov,
61, 918-930. S.I. Novel antifungal defensins from Nigella sativa L. seeds. Plant
[16] Egorov, T.A.; Odintsova, T.I.; Pukhalsky, V.A.; Grishin, E.V. Physiol. Biochem., 2011, 49, 131-137.
Diversity of wheat antimicrobial peptides. Peptides, 2005, 26, [38] Osborn, R.W.; De Samblanx, G.W.; Thevissen, K.; Goderis, I.;
2064-2073. Torrekens, S.; Van Leuven, F.; Attenborough, S.; Rees, S.B.;
[17] Terras, F.R.G.; Eggermont, K.; Kovaleva, V.; Raikhel, N.V.; Os- Broekaert, W.F. Isolation and characterization of plant defensins
born, R.W.; Kester, A.; Rees, S.B.; Torrekens, S.; Van Leuven, F.; from seeds of Asteraceae, Fabaceae, Hippocastanaceae and Saxi-
Vanderleyden, J.; Cammue, B.P.A.; Broekaert, W.F. Small cystein- fragaceae. FEBS Lett., 1995, 368, 257-262.
rich antifungal proteins from radish: their role in host defense. [39] Chen, G.H.; Hsu, M.P.; Tan, C.H.; Sung, H.Y.; Kuo, C.G.; Fan,
Plant Cell, 1995, 7, 573-588. M.J.; Chen, H.M.; Chen, S.; Chen, C.S. Cloning and characteriza-
[18] Giuliani, A.; Pirri, G.; Nicoletto, S.F. Antimicrobial peptides: an tion of a plant defensin VaD1 from azuki bean. J. Agric. Food
overview of a promising class of therapeutics. Cent. Eur. J. Biol., Chem., 2005, 53, 982-988.
2007, 2(1), 1-33. [40] Finkina, E.I.; Shramova, E.I.; Tagaev, A.A.; Ovchinnikova, T.V. A
[19] Fischbach, M.A.; Walsh, C.T. Antibiotics for emerging pathogens. novel defensin from the lentil Lens culinaris seeds. Biochem. Bio-
Science, 2009, 325, 1089-1093. phys. Res. Commun., 2008, 371, 860-865.
[20] Nelson, R. Antibiotic development pipeline runs dry. Lancet, 2003, [41] Gerlach, S.L.; Burman, R.; Bohlin, L.; Mondal, D.; Göransson, U.
362, 1726-1727. Isolation, characterization, and bioactivity of cyclotides from the
[21] Gudlaugsson, O.; Gillespie, S.; Lee, K.; Vande, B.J.; Hu, J; Messer, micronesian plant Psychotria Leptothyrsa. J. Nat. Prod., 2010, 73,
S.; Herwaldt, L.; Pfaller, M.; Diekema, D. Attributable mortality 1207-1213.
of nosocomial candidemia, revisited. Clin. Infect. Dis., 2003, 37, [42] Fant, F.; Vranken, W.; Broekaert, W.; Borremana, F. Determina-
1172-1177. tion of the three-dimensional solution structure of Raphanus sativus
[22] Thevissen, K.; Kristensen, H.H.; Thomma, B.P.H.J.; Cammue, antifungal protein by HNMR. J. Mol. Biol., 1998, 279, 257-270.
B.P.A; François, I.E.J.A. Therapeutic potential of antifungal plant [43] Almeida, M.S.; Cabral, K.M.S.; Kurtenbach, E.; Almeida, F.C.L.;
and insect defensins. Drug Discov. Today, 2007, 12(21-22), 996- Valente, A.P. Solution structure of Pisum sativum defensin 1 by
971. high resolution NMR: plant defensins, identical backbone with dif-
[23] Santos, I.S.; Da Cunha, M.; Machado, O.L.T.; Gomes, V.M. A ferent mechanisms of action. J. Mol. Biol., 2002, 215, 403-410.
chitinase from Adenanthera pavonina L. seeds: purification, char- [44] Thomma, B.P.H.J.; Cammue, B.P.A.; Thevissen, K. Plant defens-
acterisation and immunolocalisation. Plant Sci., 2004, 167, 1203- ins. Planta, 2002, 216, 193-202
1210 [45] Lay, F.T.; Brugliera, F.; Anderson, M.A. Isolation and properties of
[24] Santos, I.S.; Oliveira, A.E.A.; Cunha, M.; Machado, O.L.T.; Ne- floral defensins from ornamental tobacco and petunia. Plant
ves-Ferreira, A.G.C.; Fernandes, K.V.S.; Carvalho, A.O.; Perales, Physiol., 2003, 131, 1283-1293.
J.; Gomes, V.M. Expression of chitinase in Adenanthera pavonina [46] Terras, F.R.G.; Torrekens, S.; van Leuven, F.; Osborn, R.W.; Van-
seedlings. Physiol. Plant., 2007, 131, 80-88. deleyden, J.; Cammue, B.P.A.; Broekaert, W.F. A new family of
[25] Terras, F.R.G.; Schoofs, H.M.E.; De Bolle, M.F.C.; Van Leuven, basic cysteine-rich plant antifungal proteins from Brassicaceae spe-
F.; Rees, S.B.; Vanderleyden, J.; Cammue, B.P.A.; Broekaert, W.F. cies. FEBS Lett., 1993, 316(3), 233-240.
Analysis of two novel classes of plant antifungal proteins from rad- [47] Suter, B.; Auerbach, D.; Stagljar, I. Yeast-based functional genom-
ish (Raphanus sativus L.) seeds. J. Biol. Chem., 1992, 267, 15301- ics and proteomics technologies: the first 15 years and beyond. Bio-
15309. techniques, 2006, 40, 625-644.
[26] Bradford, M.M. A rapid and sensitive method for the quantitation [48] Shirtliff, M.E.; Krom, B.P.; Meijering, R.A.M.; Peters, B.M.; Zhu,
of microgram quantities of protein utilizing the principle of dye J.; Scheper, M.A.; Harris, M.L.; Jabra-Rizk, M.A. Farnesol-
binding. Anal. Biochem., 1976, 72, 248-254. induced apoptosis in Candida albicans. Antimicrob. Agents Che-
[27] Schägger, H.; Von Jagow, G.; Tricine-sodium dodecylsulfate poly- mother., 2009, 2392-2401.
acrylamide gel electrophoresis for the separation of proteins in the [49] Thevissen, K.; François, I.E.J.A.; Takemoto, J.Y.; Ferket, K.K.A.;
range from 1 to 100 kDa. Anal. Biochem., 1987, 166, 368-379. Meert, E.M.K.; Cammue, B.P.A. DmAPM1, an antifungal plant de-
[28] Towbin, H.; Stachelin, N.T.; Gordon, J. Electrophoretic transfer of fensin from dahlia (Dahlia merckii), interacts with sphingolidids
proteins from polyacrylamide gels to nitrocellulose sheets; proce- from Saccharomyces cerevisiae. FEMS Microbiol. Lett., 2003, 226,
dures and some applications. Proc. Natl. Acad. Sci., 1979, 76, 169-173.
4350-4354. [50] Thevissen, K.; Warnecke, D.C.; François, I.E.J.A.; Leipelt, M.;
[29] Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Heinz, E.; Ott, C.; Zähringer, U.; Thomma, B.P.H.J.; Ferket,
Basic local alignment search tool. J. Mol. Biol., 1990, 215, 403- K.K.A.; Cammue, B.P.A. Defensins from insects and plants inter-
410. act with fungal glucosylceramides. J. Biol. Chem., 2004, 279(6),
[30] Thompson, J.D.; Higgins, D.G.; Gibson, T.J. CLUSTAL W: im- 3900-3905.
proving the sensitivity of progressive multiple sequence alignment [51] Dickson, R.C.; Lester, R.L. Yeast sphingolipids. Biochim. Biophys.
through sequence weighting, position-specific gap penalties and Acta, 1999, 1426, 347-357.
weight matrix choice. Nucleic Acids Res., 1994, 22, 4673-80. [52] Warnecke, D.; Heinz, E. Recently discovered functions of gluco-
[31] Broekaert, W.F.; Terras, F.R.G.; Cammue, B.P.A.; Vanderleyden, sylceramides in plants and fungi. Cell. Mol. Life Sci., 2003, 60,
J. An automated quantitative assay for fungal growth inhibition. 919-941.
FEMS Microbiol. Lett., 1990, 69, 55-60. [53] Van der Weerden, N.L.; Lay, F.T.; Anderson, M.A. The plant
[32] Thevissen, K.; Terras, F.R.G.; Broekaert, W.F. Permeabilization of defensin, NaD1, enters the cytoplasm of Fusarium oxysporum hy-
fungal membranes by plant defensins inhibits fungal growth. Appl. phae. J. Biol. Chem., 2008, 283, 14445-14452.
Environ. Microbiol., 1999, 65(12), 5451-5458. [54] Mangoni, M.L.; Saugar, J.M.; Dellisanti, M.; Barra, D.; Simmaco,
[33] Wijaya, R.; Neumann, G.M.; Condron, R.; Hughes, A.B.; Poly, M.; Rivas, L. Temporins, small antimicrobial peptides with
G.M. Defense proteins from seed of Cassia fistula include a lipid leishmanicidal activity. J. Biol. Chem., 2005, 280, 984-990.
transfer protein homologue and a protease inhibitory plant de- [55] Regente, M.C.; Giudici, A.M.; Villalaín, J.; de la Canal, L. The
fensin. Plant Sci., 2000, 159, 243-255. cytotoxic properties of a plant lipid transfer protein involve mem-
Antimicrobial Peptides from Adenanthera pavonina Protein & Peptide Letters, 2012, Vol. 19, No. 5 529

brane permeabilization of target cells. Lett. Appl. Microbiol., 2005, [59] Steffen, H.; Rieg, S.; Wiedemann, I.; Kalbacher, H.; Deeg, M.;
40, 183-189. Sahl, H.G.; Peschel, A.; Götz, F.; Garbe, C.; Schittek, B. Naturally
[56] Hwang, B.; Hwang, J.S.; Lee, J.; Lee, D.G. Antifungal properties processed dermcidin-derived peptides do not permeabilize bacterial
and mode of action of psacotheasin, a novel knottin-type peptide membranes and kill microorganisms irrespective of their charge.
derived from Psacothea hilaris. Biochem. Biophys. Res. Commun., Antimicrob. Agents Chemother., 2006, 50(8), 2608-2620.
2010, 400(3), 352-357. [60] Aerts, A.M.; Carmona-Gutierrez, D.; Lefevre, S.; Govaert, G.;
[57] Vylkova, S.; Nayyar, N.; Li, W.; Edgerton, M. Human -defensins François, I.E.J.A.; Madeo, F.; Santos, R.; Cammue, B.P.A.; The-
kill Candida albicans in an energy-dependent and salt-sensitive vissen, K. The antifungal plant defensin RsAFP2 from radish in-
manner without causing membrane disruption. Antimicrob. Agents duces apoptosis in a metacaspase independent way in Candida al-
Chemother., 2007, 51(1), 154-161. bicans. FEBS Lett., 2009, 583, 2513-2516.
[58] Koshlukova, S.E.; Lloyd, T.L. Araujo, M.W.; Edgerton, M. Sali- [61] Aerts, A.M.; François, I.E.J.A.; Meert, E.M.K.; Li, Q.T.; Cammue,
vary histatin 5 induces non-lytic release of ATP from Candida al- B.P.A.; Thevissen, K. The antifungal activity of RsAFP2, a plant
bicans leading to cell death. J. Biol. Chem., 1999, 274, 18872- defensin from Raphanus sativus, involves the induction of reactive
18879. oxygen species in Candida albicans. J. Mol. Microbiol. Biotech-
nol., 2007, 13, 243-247.

Received: February 1, 2011 Revised: November 22, 2011 Accepted: November 22, 2011

You might also like