ArticlePDF Available

Purification and biochemical characterization of a serine proteinase inhibitor from Derris trifoliata Lour. seeds: Insight into structural and antimalarial features

Authors:

Abstract and Figures

A potent serine proteinase inhibitor was isolated and characterized from the seeds of the tropical legume liana, Derris trifoliata (DtTCI) by ammonium sulfate precipitation, ion exchange chromatography and gel filtration chromatography. SDS-PAGE as well as MALDI-TOF analysis showed that DtTCI is a single polypeptide chain with a molecular mass of approximately 20 kDa. DtTCI has three isoinhibitors (pI: 4.55, 5.34 and 5.72) and, inhibited both trypsin and chymotrypsin in a 1:1 molar ratio. Both Dixon plots and Lineweaver-Burk double reciprocal plots revealed a competitive inhibition of trypsin and chymotrypsin activity, with inhibition constants (K(i)) of 1.7x10(-10) and 1.25x10(-10) M, respectively. N-terminal sequence of DtTCI showed over 50% similarity with numerous Kunitz-type inhibitors of the Papilionoideae subfamily. High pH amplitude and broad temperature optima were noted for DtTCI, and time course experiments indicated a gradual loss in inhibitory potency on treatment with dithiothreitol (DTT). Circular Dichroism (CD) spectrum of native DtTCI revealed an unordered structure whereas exposure to thermal-pH extremes, DTT and guanidine hydrochloride (Gdn HCl) suggested that an abundance of beta-sheets along with intramolecular disulfide bonds provide conformational stability to the active site of DtTCI, and that severity of denaturants cause structural modifications promoting inhibitory inactivity. Antimalarial studies of DtTCI indicate it to be a potent antiparasitic agent.
Content may be subject to copyright.
This article appeared in a journal published by Elsevier. The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy
Purification and biochemical characterization of a serine proteinase inhibitor from
Derris trifoliata Lour. seeds: Insight into structural and antimalarial features
Arindam Bhattacharyya
a,b,*
, Cherukuri R. Babu
a
a
Centre for Environmental Management of Degraded Ecosystems, School of Environmental Studies, University of Delhi, Delhi 110 007, India
b
Malaria Research Group, International Centre for Genetic Engineering and Biotechnology (ICGEB), Aruna Asaf Ali Marg, New Delhi 110 067, India
article info
Article history:
Received 31 March 2009
Accepted 2 April 2009
Available online 4 May 2009
Keywords:
Derris trifoliata
Papilionoideae
Purification
Characterization
Conformation
Serine proteinase inhibitor
Circular Dichroism (CD)
Antimalarial
abstract
A potent serine proteinase inhibitor was isolated and characterized from the seeds of the tropical legume
liana, Derris trifoliata (DtTCI) by ammonium sulfate precipitation, ion exchange chromatography and gel
filtration chromatography. SDS–PAGE as well as MALDI-TOF analysis showed that DtTCI is a single poly-
peptide chain with a molecular mass of 20 kDa. DtTCI has three isoinhibitors (pI: 4.55, 5.34 and 5.72)
and, inhibited both trypsin and chymotrypsin in a 1:1 molar ratio. Both Dixon plots and Lineweaver–Burk
double reciprocal plots revealed a competitive inhibition of trypsin and chymotrypsin activity, with inhi-
bition constants (K
i
) of 1.7 10
10
and 1.25 10
10
M, respectively. N-terminal sequence of DtTCI
showed over 50% similarity with numerous Kunitz-type inhibitors of the Papilionoideae subfamily. High
pH amplitude and broad temperature optima were noted for DtTCI, and time course experiments indi-
cated a gradual loss in inhibitory potency on treatment with dithiothreitol (DTT). Circular Dichroism
(CD) spectrum of native DtTCI revealed an unordered structure whereas exposure to thermal-pH
extremes, DTT and guanidine hydrochloride (Gdn HCl) suggested that an abundance of b-sheets along
with intramolecular disulfide bonds provide conformational stability to the active site of DtTCI, and that
severity of denaturants cause structural modifications promoting inhibitory inactivity. Antimalarial stud-
ies of DtTCI indicate it to be a potent antiparasitic agent.
Ó2009 Elsevier Ltd. All rights reserved.
1. Introduction
Tropical rain forest legumes represent a major repository of
proteinase inhibitor (PIs) diversity, and their co-evolutionary
trends with the insect proteinases ensure a dynamic duel among
them, at the molecular level. Although PIs are believed to form a
wide-spectrum defense mechanism, they are involved in diverse
biochemical functions such as in regulation of metabolism/proteo-
lytic cascades (Shewry and Lucas, 1997), suppression of in vitro and
in vivo replication of retroviruses (Gustafson et al., 1994), hemo-
lytic activity (Tam et al., 1999), antifungal property (Swartz et al.,
1977), as anticarcinogenic agents (Kennedy, 1998) and inhibition
of intraerythrocytic development of Plasmodium falciparum (Rock-
ett et al., 1990).
PIs vary widely in their botanical origin, and in their cause and
consequence among the reproductive organs, storage organs and
vegetative tissues of most plant families. In higher plants, several
gene families have been characterized, particularly those constitut-
ing the serine protease inhibitors from Leguminosae, Solanaceae
and Graminae. Earlier studies on the distribution pattern of PIs
among seeds of leguminous trees showed an evolutionary relation-
ship between the PI family and the legume sub-families (Norioka
et al., 1988). Currently, 59 distinct families of PIs have been recog-
nized. On the basis of amino-acid sequences, localization of the
reactive sites, disulfide bridge topology, mechanism of action,
three-dimensional structure, and stability to heat and denaturing
agent, they have been placed under seven distinct families: Kunitz,
Bowman–Birk, Potato I, Potato II, Squash, Cereal and Mustard (De
Leo et al., 2002; Birk, 2003).
The two best characterized families of plant serine PIs are the
Kunitz-type and Bowman–Birk type inhibitors. Most serine PIs
are low-molecular mass molecules (3–25 kDa) that inhibit trypsin
and/or chymotrypsin. These families differ from each other in
mass, cysteine content and number of reactive sites (Richardson,
1977). Kunitz-type inhibitors are proteins of Mr 20 kDa, with
low cysteine content and a single reactive site, whereas the Bow-
man–Birk type inhibitors have Mr 8–10 kDa as well as high cys-
teine content and two reactive sites (Richardson, 1991).
Structurally, both Kunitz and Bowman–Birk inhibitors lack
a
-helix,
but vary in their mode of stability. Kunitz inhibitors are stabilized
chiefly by hydrophobic interactions of short stretches of hydrogen
bonded sheets (soybean Kunitz trypsin inhibitor) whereas the
disulfide linkages in the Bowman–Birk inhibitors minimize their
0031-9422/$ - see front matter Ó2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.phytochem.2009.04.001
*Corresponding author. Present Address: Department of Science and Technology,
Ministry of Science and Technology, Technology Bhawan, New Mehrauli Road, New
Delhi 110 016, India. Tel./fax: +91 011 26590539.
E-mail address: ady1111@rediffmail.com (A. Bhattacharyya).
Phytochemistry 70 (2009) 703–712
Contents lists available at ScienceDirect
Phytochemistry
journal homepage: www.elsevier.com/locate/phytochem
Author's personal copy
conformational entropy and enhance their stability (Sweet et al.,
1974; Ramasarma et al., 1995).
The complex biology and life cycle of the apicomplexan parasite
P.falciparum has hindered attempts to control malarial infections
and also prevent transmission. Several attempts toward an effec-
tive vaccine system as well as antimalarial drugs have had only
partial success, and demands for new therapeutics have always ex-
ceeded the available treatments due to massive antigenic variation
and rapid drug resistive prowess of the parasite. In spite of these
limitations, regular updates on the parasite genome biology pro-
vide a platform to renew our efforts towards complete control of
malarial infection. In fact, a recent survey of the parasite genome
has revealed that over 90 putative proteases belonging to 26 fam-
ilies of five clans constitute the parasite genome (Wu et al., 1992).
These proteases (11% aspartic, 36% cysteine, 22% metallo, 17% ser-
ine and 14% threonine) are engaged in a functionally diverse but
well orchestrated cascade of proteolytic processes throughout its
life cycle with particular emphasis in host cell invasion, nutrition
and growth, and the processing of precursor proteins (Klemba
and Goldberg, 2002; Pattanaik et al., 2003). Thus, a renewed ap-
proach towards the discovery of new antiparasitic molecules has
been the design and discovery of different types of PIs towards
these parasite proteases, as PIs of different classes have been
shown to arrest the parasite invasion of erythrocyte and prevent
its survival inside the host (Blackman, 2000; Rosenthal, 2002).
Co-evolutionary trade-offs, which are the precursors of a pleth-
ora of plant–herbivore interactions in tropical forests, engenders a
vast reservoir of PIs in leguminous plants, some of which may
prove to be effective antimalarial chemotherapeutics. Moreover,
not much attempt has been done on plant derived PIs as antimal-
arials and even the limited studies conducted (Dejkriengkraikhul
and Wilairat, 1983) have yielded only partial success. Thus, in light
of the aforesaid facts, we report here the purification, structure–
function characterization and antiplasmodial property of a serine
PI from the seeds of Derris trifoliata (DtTCI), a commonly occurring
liana along mangrove and coastal forests of Great Nicobar Island,
Andaman and Nicobar Islands, India.
2. Result and discussion
2.1. Purification of the Derris seed proteinase inhibitor
A crude inhibitor (CI) preparation was obtained by saline
extraction of ground seeds of D.trifoliata showing 53% and 58%
inhibition of trypsin and chymotrypsin activity, respectively.
Table 1 demonstrates the inhibitor characteristics after (NH
4
)
2
SO
4
precipitation, DEAE 52 cellulose ion exchange chromatography, gel
filtration chromatography on Superdex S-75 and Q-Sepharose ion
exchange chromatography. DEAE ion exchange chromatography
of the 0–75% (NH
4
)
2
SO
4
precipitated inhibitor revealed four major
protein peaks, each possessing either antitrypsin or antichymo-
trypsin activity (Fig. 1a). The protein eluted with 0.2 M NaCl
(fractions: 82–100 ml) demonstrated a strong trypsin inhibitory
(11.9 TIU/A
280
) as well as chymotrypsin inhibitory (12.2
CIU/ A
280
) activity (Table 1), and was subsequently applied on a
Superdex S-75 column. The elution pattern showed a major leading
peak (DtTCI-1) followed by a minor trailing peak (DtTCI-2)
(Fig. 1b). Expectedly, DtTCI-1 inactivated over 90% of both trypsin
and chymotrypsin activity, with similar specific activities towards
both the serine proteases. DtTCI-1 was rechromatographed on
Q-Sepharose which further resolved into two main peaks (P2 and
P3) (Fig. 1c). Purified DtTCI (P2) represented 0.1% of the total seed
proteins of D.trifoliata. These results concur well with previously
studied Papilionoid inhibitors (Haq and Khan, 2003; Garcia et al.,
2004).
2.2. Molecular mass evaluation and isoinhibitor forms
Analyses of purified DtTCI by SDS–PAGE gel electrophoresis
both in presence as well as absence of DTT yielded a single poly-
peptide chain corresponding to a molecular weight of 20 kDa
(Fig. 1c). In addition, MALDI-TOF analyses also showed a molecular
mass of 20.1 kDa (Fig. 2a). Studies conducted previously on Kunitz
inhibitors belonging Papilionoideae sub-family (Garcia et al., 2004;
Gomes et al., 2005), have also reported similar results. Isoelectric
focusing of DtTCI revealed the acidic nature of the inhibitor and,
also the presence of three isoinhibitors (Fig. 2b). The three isoin-
hibitors had pIof 4.55, 5.34 and 5.72, respectively. Similar isoinhib-
itor forms and a strong acidic trait are typical attributes of several
Kunitz PIs (Richardson, 1991). Physiologically, occurrence of such
isoinhibitors may be a part of the survival strategies evolved by
the host plant.
2.3. N-terminal amino acid sequence analysis
The N-terminal sequence analysis of DtTCI exhibited a strong
similarity with other members of the Kunitz PI from the Papilionoi-
deae subfamily (Onesti et al., 1991; Kouzuma et al., 1992; Terada
et al., 1994; Datta et al., 2001). In fact, DtTCI showed over 70% se-
quence similarity with PIs of Erythrina variegata and Canavalia line-
ata and close resemblance to trypsin inhibitors from Psophocarpus
tetragonolobus and Erythrina caffra (Fig. 3a).
Table 1
Summary of the purification of a serine proteinase inhibitor from Derris trifoliata (DtTCI).
Step/characteristics Total protein (mg) % Inhibition Specific activity % Recovery Purification factor
T
a
C
b
T(TIU/ A
280
)
c
C(CIU/ A
280
)
d
TC
Crude inhibitor 632.5 53.3 57.8 1.1 1.3 100 1.0 1.0
Ammonium sulfate precipitation 0–75% 161.0 75.1 68.3 3.0 3.6 25.5 2.7 2.8
DEAE 52 cellulose chromatography NaCl (M) 0.1 33.3 76.7 87.5 5.2 5.5 5.3 4.7 4.2
0.2 26.8 82.2 89.2 11.9 12.2 4.2 10.8 9.4
0.3 15.8 78.4 70.1 10.2 10 2.5 9.1 7.7
0.4 7.1 53.8 31.6 9.2 2.6 1.1 8.4 2.0
Superdex S-75 gel filtration 6.3 92.3 97.6 56.5 58.6 0. 7 51.4 45.1
Q sepharose chromatography NaCl (M) 0.2 0.8 94.7 98.2 70.8 71.1 0.1 64.4 54.7
0.3 1.5 79.6 82.5 12.5 11.8 0.2 11.4 9.1
Note: One trypsin or chymotrypsin unit is defined as 1
l
mol of substrate hydrolyzed per minute of reaction. One inhibition unit is defined as unit of enzyme inhibited. Specific
activity is defined as inhibition units per absorbance unit, at 280 nm, of the inhibitor.
a
Trypsin.
b
Chymotrypsin.
c
Trypsin inhibitory units.
d
Chymotrypsin inhibitory units.
704 A. Bhattacharyya, C.R. Babu / Phytochemistry 70 (2009) 703–712
Author's personal copy
2.4. Inhibitor kinetics and inhibition constant
DtTCI suppressed both tryptic as well as chymotryptic activity
with similar potency. The titration profiles for both trypsin and
chymotrypsin remained linear up to 90% inhibition and extrapo-
lation of the curve engendered a 1:1 trypsin/chymotrypsin–DtTCI
complex (Fig. 3b).
Predictably, kinetic studies on the initial rates of reaction in the
presence or absence of DtTCI followed the Michaelis–Menten
equation, and both the Lineweaver–Burk double reciprocal plots
(a)
(b)
(c)
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
1 4 7 10 13 16 19 22 25 28 31 34 37 40 43 46 49 52 55 58 61 64
Fraction number
Absorbance at 280 nm
0
10
20
30
40
50
60
70
80
90
100
% inhibition
P3
P2
P1
NaCl (M)
0.6
0.5
0.4
0.3
0.2
0.1
66.0
20.1
14.4
M P2 P3
R NR
1 2 3 4
Fig. 1. (a) Ion exchange chromatography of the 0–75% ammonium sulfate fractionated proteins of D.trifoliata seed extract, on DE 52-Cellulose. Column dimensions:
2.5 18 cm; Starting buffer: 20 mM Tris HCl (pH 8). The line cutting across the chromatogram indicates application of a step gradient in the NaCl concentration in the starting
buffer. (b) Size-exclusion chromatography of 0.2 M NaCl DtTCI fraction (eluted from the DE-52 Cellulose column) onto a Superdex S-75 column. Inset: SDS–PAGE profile of
DtTCI-1. (c) Elution profile of DtTCI-1 on Q-Sepharose chromatography column (2.5 17 cm) equilibrated with 20 mM Tris HCl (pH 8). Protein elution was performed using a
linear gradient (0.1–0.5 M) of NaCl. Fractions (3 mL) were collected and assayed for antitryptic (
D
) and antichymotryptic (h) activities. Inset: SDS–PAGE of 0.2 M (lanes 2 and
3: under reducing (R) and non-reducing (NR) conditions, respectively) and 0.3 M (lane 4) NaCl fractions of DtTCI following re-ion exchange chromatography on Q-Sepharose.
Lane 1: protein molecular weight marker.
A. Bhattacharyya, C.R. Babu / Phytochemistry 70 (2009) 703–712 705
Author's personal copy
(Fig. 4a and c) and Dixon plots (Fig. 4b and d) indicated a compet-
itive inhibition of both trypsin and chymotrypsin activity, respec-
tively. The inhibition constants (K
i
) of DtTCI against both trypsin
(TAME hydrolyzing units) and chymotrypsin (BTEE hydrolyzing
units) activity were 1.7 10
10
and 1.25 10
10
M, respec-
tively (Fig. 4b and d). Similar K
i
values were also recorded for PIs
isolated from other legumes (Chaudhary et al., 2008; Bhattachar-
yya et al., 2006; Garcia et al., 2004).
2.5. Stability attributes
DtTCI showed 80% and over 60% inhibition of trypsin/chymo-
trypsin activity on a temperature gradient of 15–75 °C and over a
pH range of 2–12 (data not shown), respectively. Interestingly,
DtTCI retained over a 2-fold higher trypsin/chymotrypsin inhibi-
tion following exposure to 6 N HCl as compared to conditions of
1 N NaOH (Table 2). Earlier studies on many of the Kunitz PIs have
also reported similar defiance to extremes of thermo-pH condi-
tions and have attributed this stability to the apparent ability of
the inhibitor to reversible denaturation through a transient inter-
mediate, presence of aromatic (tyrosine) residues generally impli-
cated in energy transfer and disulfide bonds (Roychaudhuri et al.,
2003; Liao et al., 2007; Yoshizaki et al., 2007; Zhou et al., 2008;
Chaudhary et al., 2008). The effect of DTT on the inhibitory activity
of DtTCI was also examined. DtTCI was also able to withstand
reducing conditions of 10 mM DTT for up to 2 h following which
there was a steady loss of inhibitory activity with less than 10%
inhibitory activity at the end of 5 h (Table 2;Fig. 5a). Similar effect
of DTT was also observed for other Kunitz PIs (Garcia et al., 2004;
Lingaraju and Gowda, 2008; Zhou et al., 2008). Although intramo-
lecular disulfide linkages have been ascribed to be the prime factor
behind the functional stability of the inhibitor against denaturants
(Zhou et al., 2008; Liao et al., 2007; Macedo et al., 2007; Oliveira
et al., 2007), additional factors such as placement/arrangement of
the disulfide linkages (buried, rather than surface exposed) within
the molecule are also critical for the overall stability of the inhibi-
tor (Tetenbaum and Miller, 2001; Yoshizaki et al., 2007; Macedo
et al., 2007).
2.6. CD based conformation study
Estimation of the secondary structure of native DtTCI as well as
DtTCI exposed to extreme heat, pH, 6 N guanidine hydrochloride
and DTT in solution was carried out using near and far-ultraviolet
CD based studies. The far-UV CD spectrum of native DtTCI (Fig. 5b)
was similar to that of b-II proteins, exhibiting a sharp minimum
ellipticity at 200 nm and a positive band at 192 nm with a minor
shoulder close to 230 nm. The detailed structural composition of
the protein showed a large fraction (40%) of b-sheets (Table 2).
Similar CD spectrums were also noted for several other legume
PIs (Liao et al., 2007; Yoshizaki et al., 2007; Bhattacharyya and
Babu, 2006; Araujo et al., 2005).
CD-based conformational evaluation of DtTCI following expo-
sure to severe heat and acid–alkali conditions showed gross devi-
ations among the
a
-helix and b-sheets content which were in
concurrence with the reduced inhibitory activity (Table 2; spec-
trum not shown). The results are in agreement with experimental
studies carried out on other Kunitz PIs (Liao et al., 2007; Yoshizaki
(b)
8.65
8.45
8.15
7.35
6.85
6.55
5.85
5.20
4.55
3.50
DtTCI 1 (5.72)
DtTCI 2 (5.34)
DtTCI 3 (4.55)
(a)
Fig. 2. (a) MALDI-TOF mass spectrum of DtTCI. (b) Isoelectric focusing of DtTCI on
Ampholine polyacrylamide gel plates (pH range 3.5–9.5, Pharmacia) over a broad-
range (pI: 3–10) calibration kit. Lane 1: marker and lane 2: purified DtTCI showing
three isoinhibitors with pIat 5.72, 5.34 and 4.55.
(a)
(b)
0
10
20
30
40
50
60
70
80
90
100
00.511.52
Molar ratio (inhibitor: enzyme )
(%) Residual enzyme activity
Tr yp s in
Chymotryps in
% Alignment
Score
DtTCI ELVLDVNGDQVRNGGTYYL
EvTI ELV-DVEGEDVVNGGTYYM
ClTI --VLDTDGDMVRNGGIYYI
PtTI EELVDVEGKTVRNGGTYYL
EcTI -VLLDGNGEVVQNGGTYYL
BvTI EIVLDQNGNPVRNSGGRYY
STI DFVLDNEGNPLSNGGTYYI
Sporamin A TPVLDINGDEVRA-GGNYY
AcTI -KELLDADGDILRNGGAYY
CaTI EQVLDTNGNPLIPGDEYYI
ElTI -VLLDGNGEVVQN-GGIYY
PpTI -APLEDSLAAKLNNGTSYY
72
70
68
66
57
52
52
52
47
44
31
Fig. 3. (a) N-terminal sequence alignment of related Kunitz Proteinase Inhibitors
from Erythryna variegata,Canavalia lineata,Psophocarpus tetragonolobus,Erythryna
caffra,Bauhinia variegata,Glycine max,Ipomoea batatas,Acacia confusa,Cicer
arietinum,Erythryna latissima,Poecilanthe parviflora. (b) Titration curves of bovine
trypsin and
a
-chymotrypsin inhibition by DtTCI. Following an incubation of 5 min,
percent residual trypsin (——) and chymotrypsin (—j—) activities were measured
at 25 °C (at pH 8.1) using substrates TAME and BTEE, respectively, in the presence of
varying concentrations of DtTCI.
706 A. Bhattacharyya, C.R. Babu / Phytochemistry 70 (2009) 703–712
Author's personal copy
et al., 2007; Garcia et al., 2004; Haq and Khan, 2003; Macedo et al.,
2003). To examine the changes accorded by DtTCI at a structural
level following DTT exposure, far-UV CD measurements were per-
formed. Noticeably, marked differences in the far-UV spectrum of
native and reduced DtTCI were recorded which included a loss in
the negative band at 200 nm in the latter (Fig. 5c). Also, a minor
reduction in ellipticity of the 230 nm band was noted, which might
have been due to the chirality of the cysteine bonds. Table 2 shows
a loss in the b-sheets and a concomitant increase in the helical con-
tent of the inhibitor. Conformational changes in DtTCI due to disul-
fide bond reduction disrupted the protein stability, and partly
affected the reactive site, as evidenced by the retention of 20%
trypsin and 40% chymotrypsin inhibitory activity, respectively
after 3 h of reduction by 10 mM DTT (Fig. 5a). Similar partial pro-
tection of the active site has also been observed for a Kunitz PI from
Entada scandens (Lingaraju and Gowda, 2008). It is interesting to
note, that under extreme DTT conditions the magnitude of loss in
inhibitory activity of Kunitz PIs varies from species to species.
For instance, upon exposure to 100 mM DTT, the inhibitory activi-
ties of some of the Kunitz PIs were abolished (Garcia et al., 2004;
Macedo et al., 2003) whereas some others remained completely
unaffected (Chaudhary et al., 2008; Lehle et al., 1996). It is pre-
sumed that the intramolecular disulfide bonds and their distance
from the reactive site residues determine the functional stability
of these inhibitors in presence of denaturing agents (Garcia et al.,
2004; Lehle et al., 1996). However, studies on the susceptible (sin-
gle polypeptide) soybean trypsin inhibitor (Tetenbaum and Miller,
2001) or stable Bauhinia species inhibitors (devoid of any disulfide
linkages), indicated that intramolecular disulfide bonds in Kunitz-
type PIs may be essential, but not exclusive to the stabilization of
the reactive site loop.
Upon exposure to 6 N Gdn HCl for 2 h at 70 °C, DtTCI showed
gross conformational perturbations due to a total denaturation of
the protein (Fig. 6a and 6b). In fact, the near UV spectrum recorded
a significant decrease in amplitude for the broad negative elliptic-
ity band between 290 and 250 nm in comparison to the native
(a) (b)
(c) (d)
0
5
10
15
20
25
30
35
40
45
50
-3 -2 -1 0 1 2 3 4 5 6 7 8 9 10 11
DtTCI [x10-1 nM]
DtTCI [x10-1 nM]
BTEE (0.535 mM)
BTEE (1.07 mM)
0
3
6
9
12
15
18
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5
TAME [0.005 mM]
TAME [0.01 mM]
0
3
6
9
12
15
18
-200 -100 0 100 200
1/ [s] (TAME) mM
-1
1/ v (OD
247
mM/min /mL)
-1
1/ v (OD
247
mM/min /mL)
-1
[I] = 0 nM
[I] = 5 nM
[I] = 6.6 nM
[I] = 10 nM
Ki
Ki
0
4
8
12
16
20
24
-2 -1 0 1 2
1/ [s] (BTEE) mM) -1
1/v (OD
256
mM/min/m L)
-1
1/v (OD
256
mM/min/mL)
-1
[I] = 10 nM
[I] = 6.6 nM
[I] = 5 nM
[I] = 0 nM
Fig. 4. Kinetic analysis of trypsin and chymotrypsin inhibition by purified DtTCI. A Lineweaver–Burk plot showing competitive nature of trypsin (a) and chymotrypsin (c)
inhibition by DtTCI and Dixon plot for the determination of the inhibition constant (K
i
) of DtTCI towards trypsin (b) and chymotrypsin (d) at two different concentrations of
TAME and BTEE, respectively. The reciprocals of velocity were plotted against variable DtTCI concentration.
Table 2
CD secondary structure content and serine proteinase inhibitory activity of native and treated DtTCI. All experiments were performed in 10 mM Tris buffer/50 mM NaCl, pH 8, at
25 °C.
DtTCI (D) CD secondary structure content (%) % Inhibition
a
-Helix b-Sheet Unordered T
a
C
b
D only 4 39 57 91.3 ± 6.1 95.5 ± 6.8
D (100 °C; 2 h) 21 4 75 36.4 ± 3.2 32.6 ± 4.5
D + 6 N HCl (30 °C; 2 h) 8 30 62 72.5 ± 6.2 88.2 ± 5.7
D + 1 N NaOH (30 °C; 2 h) 19 21 60 31.4 ± 4.7 28.8 ± 4.4
D + Gdn HCl (6 N; 40 °C; 5 h) 10 19 71 19.8 ± 2.5 26.2 ± 2.8
D + Gdn HCl (6 N; 70 °C; 2 h) 23 4 73 4.2 ± 0.6 8.6 ± 0.8
D + DTT (10 mM; 37 °C; 2 h) 14 26 60 62.4 ± 5.1 73.4 ± 5.7
D + DTT (10 mM; 37 °C; 5 h) 30 8 62 5.8 ± 0.4 9.2 ± 0.6
a
Trypsin.
b
Chymotrypsin.
A. Bhattacharyya, C.R. Babu / Phytochemistry 70 (2009) 703–712 707
Author's personal copy
DtTCI (Fig. 6a). Detailed examination revealed that the total loss of
secondary structure was primarily due to a complete conversion of
the b-sheets to helical and unordered contents (Table 2). The se-
vere effect of Gdn HCl on DtTCI is in accordance with similar stud-
ies on other Kunitz PIs (Souza et al., 2000; Haq and Khan, 2003;
Leach and Fish, 1977).
2.7. Antimalarial activity
The in vitro antimalarial efficacy of DtTCI was propounded by
monitoring a decrease in the pLDH (parasite Lactate Dehydroge-
nase) activity against the two P. falciparum lines, 3D7 and FCR3.
Typically, a dose-dependent curve was observed against both the
parasite lines although marked differences in the erythrocyte inva-
sion inhibitory levels existed between them (Fig. 7a). In fact, DtTCI
showed a minor fold (1.35 times) higher inhibitory capacity
against 3D7 as compared to FCR3. The IC
50
values obtained for
DtTCI for P.falciparum 3D7 and FCR3 were 9 and 16.35
l
M, respec-
tively (Fig. 7c).
A few studies prior to our study involving the role of plant de-
rived serine PIs didn’t hold much promise (Dejkriengkraikhul and
Wilairat, 1983) and consequently, efforts to explore plant PIs as
effective antimalarial drugs were also limited. However, some re-
cent studies have advocated that proteolytic processing by serine
proteases is critical for parasite invasion of erythrocytes. For
example, Kato et al. (2005) have suggested that Apical Merozoite
Antigen 1 (AMA), an indispensable parasite protein, plays an
important but not exclusive role in invasion of human erythro-
cytes through a process that involves processing of the erythro-
cyte surface protein Kx, by a trypsin-like parasite protease.
Similarly, a recent study has also reported that secondary prote-
olytic processing of yet another vital parasite surface protein –
the Merozoite Surface Protein 7 (MSP7), is abrogated in the pres-
ence of serine protease inhibitors (Pachebat et al., 2007). Further
evidences on the role of serine proteases particularly that of chy-
motrypsin type in parasite invasion of erythrocyte have been doc-
umented by earlier studies using chymostatin as an
antiplasmodial agent (Howell et al., 2003; Dutta et al., 2003,
2005).
Our data exhibits a reduction in parasitemia following exposure
to DtTCI and this may be due to an effective blocking of the essen-
tial serine proteolytic activity of/on some major surface antigens of
the parasite. Although, several studies conducted earlier have doc-
umented PIs belonging to serine, cysteine, aspartyl and metallo-
protease classes as successful antimalarial agents (Blackman,
2000; Rosenthal, 2002), much work towards understanding the
intricate mechanisms of parasite protease–protease inhibitor
interactions still remains and need to be resolved systematically.
Moreover, ever since Wu et al. (1992) demonstrated that the gen-
ome of P.falciparum houses several classes of stage-specific prote-
ases each of which has some functional connotation, the relevance
of these molecules as effective targets for antimalarial drugs have
been recognized and renewed efforts have been galvanized in that
(a)
(b)
(c)
0
20
40
60
80
100
0 30 60 90 120 150 180 210 240 270 300
Time (Min)
% Residual inhibitory activity
% Chymotrypsin
% Trypsin
Fig. 5. Conformational stability and antiproteolytic activity of DtTCI on exposure to
DTT. Both the residual trypsin and chymotrypsin inhibitory activity were assayed by
using substrates TAME and BTEE, respectively, in 50 mM Tris–HCl, pH 8.0. All
experiments were performed in 10 mM Tris buffer/50 mM NaCl, pH 8. Each point is
the mean of three assays. (a) Time course profile of the antiproteolytic activity of
DtTCI after incubation with 10 mM DTT at 37 °C. (b) Far UV CD spectra of native
DtTCI. (c) Far-UV CD spectra of native DtTCI and DTT reduced with DtTCI (37 °C; 5 h).
Fig. 6. Effect of 6N Gdn HCl on conformational stability of DtTCI (a) Near UV CD
spectra at various temperatures (b) Far UV CD profile following treatment with 6N
Gdn HCl at 70 °C for 2 h. Inhibitor concentration of 0.5 mg/ml in 10 mM Tris buffer/
50 mM NaCl, pH 8 was used. Each point is the mean of three assays.
708 A. Bhattacharyya, C.R. Babu / Phytochemistry 70 (2009) 703–712
Author's personal copy
direction. Consequently, the role of plant serine PIs needs to be re-
explored as effective agents against the malarial parasite as well as
the spread of malarial disease.
3. Materials and methods
3.1. Plant material
Derris trifoliata Lour. seeds were manually harvested from the
coastal/mangrove forests of Great Nicobar Island, Andaman and
Nicobar Islands, India. Seeds were air-dried and stored at 20 °C
before use.
3.2. Isolation and purification of DtTCI
Fifty gram of D.trifoliata seeds were soaked overnight in 0.15 N
NaCl. The swollen cotyledons were then homogenized with 350 ml
of saline Tris buffer (20 mM Tris, pH 8.0; 0.15 M NaCl) containing
1 mM sodium metabisulfite at 4 °C. Following filtration through a
chilled 4-fold muslin cloth, the homogenate was centrifuged for
15 min at 12,000 gat 4 °C. The supernatant (crude extract) ob-
tained was fractionated by a 0–75% (NH
4
)
2
SO
4
precipitation. Sub-
sequently, the precipitated protein was dialyzed against cold
deionized water, lyophilized and submitted to proteinase inhibi-
tory assays. The dialyzed protein showing high antitryptic activity
was applied to an anion exchanger column DEAE 52, pre-equili-
brated with 20 mM Tris buffer, pH 8.0. Protein fractions (5 ml)
were eluted using a step gradient (0.1–0.5 M) of NaCl at a flow rate
of 0.5 ml/min. Fractions eluted with 0.2 M NaCl were pooled, dia-
lyzed, lyophilized and submitted (1.0 mg/ml) to a Superdex S-75
column pre-equilibrated with 0.1 M Tris–HCl buffer (pH 8.0) con-
taining 0.1 M NaCl, for gel filtration chromatography. DtTCI active
fractions (0.5 ml; flow rate: 20 ml/h) were collected. The column
fractions with DtTCI activity were dialyzed, concentrated and
loaded onto a strong anion exchanger column Q-Sepharose Fast
Flow pre-equilibrated with 50 mM Tris–HCl buffer (pH 8.0),
50 mM NaCl. The retained proteins were eluted with 50 mM
Tris–HCl (pH 8.0) containing 1 M NaCl at flow rate of 1 ml/min
and dialyzed against cold deionized water. The antitryptic peak
(P2) was pooled and concentrated for further analyses.
3.3. Proteinase inhibition studies
Following Birk (1976), trypsin and chymotrypsin inhibitory as-
says were carried out using the substrates TAME (tosyl–arginyl–
methyl ester hydrochloride) and BTEE (n-benzoyl tyrosine ethyl es-
ter) respectively. The remaining esterolytic activity of trypsin and
chymotrypsin towards their respective substrates in presence of
DtTCI were estimated. Both the proteinases were pre-incubated
with DtTCI for 5 min at 25 °C, in 1 ml Tris buffer (46 mM) contain-
ing 11.5 mM CaCl
2
, pH 8.1, prior to the reaction. One trypsin or
chymotrypsin unit is defined as 1
l
mol of substrate hydrolyzed
per minute of reaction. One inhibition unit is defined as unit of en-
zyme inhibited. Specific activity is defined as trypsin inhibition
units (TIU) per absorbance unit, at 280 nm (A
280
), of the inhibitor.
3.4. Estimation of proteins
Protein content was measured by Coomassie blue staining
according to the procedure of Bradford (1976) as well as from
the absorbance at 280 nm. BSA (1 mg/ml) was used as a protein
standard.
3.5. Reduction and S-alkylation
As outlined by Crestfield et al. (1963), DtTCI (5 mg) was reduced
with 0.1 M dithiothreitol and S-carboxymethylated using iodoace-
tic acid. Subsequent to desalting, DtTCI was redissolved in 0.05 M
Tris buffer pH 8.0, and gel filtrated on Superdex S-75, equilibrated
with the same buffer.
3.6. SDS–PAGE profiles
Following Laemmli (1970), purified DtTCI was subjected to
SDS–PAGE using 5–15% polyacrylamide slab gel system, after pre-
treating the protein sample with Laemmli’s buffer. The gels were
stained by Coomassie brilliant blue R250. A low molecular weight
range (14–116 kDa) of protein molecular weight markers (Amer-
sham Biosciences), were used.
(a)
(b)
(c)
0
10
20
30
40
50
60
70
80
Proteinase Inhibitor (
µ
M)
% Erythrocyte invasion inhibition
DtTCI
Chymostatin
DtTCI
0 0 1.3 3.42 6.17 9.83 14.9 35.4 48.2 69.5
Chymostatin
0 0 0.08 0.48 2.67 6.78 11.3 26.8 36.6 52.2
0 0.03 0.06 0.13 0.25 0.5 1 5 10 25
0
10
20
30
40
50
60
70
80
Proteinase Inhibitor (
µ
M)
% Erythrocyte invasion inhibition
DtTCI
Chymostatin
DtTCI
0 0 1.12 2.04 3.46 5.76 10.1 30 40.2 51.3
Chymostatin
0 0 0.03 0.21 0.88 2.05 4.73 17.2 26.5 33.7
0 0.03 0.06 0.13 0.25 0.5 1 5 10 25
Fig. 7. LDH based growth inhibition assay of DtTCI and chymostatin on the growth
of P.falciparum strains 3D7 (a) and FCR3 (b) Mature trophozoite enriched parasite
cultures (0.3 ± 0.1%; 2% hematocrit) were incubated with DtTCI for 42 h at 37 °C and
examined for inhibition of erythrocyte invasion. Each point represents the
mean ± S.E. from one representative experiment in triplicate. (c) In vitro efficacy
of DtTCI on inhibition of P.falciparum (3D7 and FCR3 strains) growth. GraphPad
Prism (5.1 program) was employed to generate the inhibition curve. Each point
represents mean ± SE of three LDH based GIA assays.
A. Bhattacharyya, C.R. Babu / Phytochemistry 70 (2009) 703–712 709
Author's personal copy
3.7. Mass spectrometry
Matrix-assisted laser desorption/ionization time of flight mass
spectrometry (MALDI-TOF/MS) was used to analyze the purity
and absolute molecular mass of DtTCI using a Bruker Daltonics
Ò
model Ultraflex spectrometer. Spectra were recorded in the posi-
tive-ion mode at 50 Hz laser frequency. Samples were prepared
by mixing equal volumes of sinapinic acid in 0.1% TFA (aq.), aceto-
nitrile (2:1), and the inhibitor solution. From this mixture, 1
l
l was
spotted on to the sample plate and allowed to evaporate to dry-
ness. Bovine trypsinogen was used for internal calibration.
3.8. N-terminal sequencing
The first 20 amino acids from the N-terminal end were se-
quenced by Edman degradation method on an Applied Biosystem
Procise Sequencer. Percentage sequence identity with trypsin
inhibitors from other members of Mimosoideae sub-family was
determined using BCM search launcher for multiple sequence
alignment (Worley et al., 1998) and BOXSHADE 3.21 was used
for shading of multiple alignment files.
3.9. Isoelectric focusing
Isoelectric focusing was performed using a Phast System appa-
ratus from Pharmacia LKB Biotechnology to separate the iso-inhib-
itors of DtTCI. Phast Gel IEF 3–9, which operates in the pH range 3–
9, and Ampholine polyacrylamide gel plates from Pharmacia (pH
range 3.5–9.5) were used together with a Pharmacia broad-range
pIcalibration kit containing proteins with various isoelectric points
ranging from 3 to 10.
3.10. Kinetic analysis of DtTCI
Kinetic analyses of DtTCI activity were carried out by using Dix-
on plot estimation (Segel, 1975). A series of DtTCI concentrations
(nM) of 4, 5, 6.6 and 10 were used to determine the inhibition con-
stant (K
i
). Enzyme inhibition was characterized at two different
substrate concentrations [TAME (0.005 and 0.01 mM) and BTEE
(0.53 and 1.07 mM)], for trypsin and chymotrypsin, respectively.
The initial slope
v
was determined for each inhibitor concentration.
The reciprocal velocity (1/
v
) versus [PI], for each substrate concen-
tration, [S
1
] and [S
2
], was plotted. A single regression line for each
[S] was obtained, and the K
i
was calculated from the intersection of
the two lines. The velocity of the reaction was expressed as 1/v
(OD
247
mM/min/ml)
1
. The mechanism of inhibition was deter-
mined using Lineweaver–Burk plots, in which the inverse of the
initial rate was plotted against the inverse of the substrate concen-
tration in the absence or presence of DtTCI.
3.11. Biological properties of DtTCI
The temperature optima of purified DtTCI was estimated by its
trypsin and chymotrypsin inhibitory activity, under varying tem-
peratures (5–100 °C, with an increment of 10 °C) in 0.05 M Tris–
HCl buffer, pH 8.0. For each temperature, an incubation period of
60 min duration was maintained. Both antitrypsin and antichymo-
trypsin assays under a range of pH (2–12) conditions were used to
assess the pH optima of DtTCI. Purified DtTCI was mixed with buf-
fers of different pH respectively in the ratio of 1:1 (v/v), and incu-
bated for 3 h. The buffers used were 0.02 M each of glycine–HCl
(pH 2.0–3.5), sodium phosphate (pH 6–7.5), Tris–HCl (pH 8.0–
10.0) and glycine–NaOH (pH 10.5–13.0). Three replicates were
maintained for each temperature and pH condition. In addition, ef-
fects of extreme acidic and basic conditions on the inhibitory
capacity of DtTCI were examined by treatment of the inhibitor
with 6 N HCl and 1 N NaOH. Further, DtTCI was (0.5 mg/ml) was
incubated with dithiothreitol (DTT) at final concentrations of
2 mM and 10 mM for 30–300 min each at 37 °C. After the stipu-
lated time period the reaction was terminated using 4 mM of iodo-
acetamide, and antitryptic activity of DtTCI was assessed.
Replicates in triplicate were maintained for the different chemical
treatments.
3.12. Circular Dichroism (CD) of DtTCI
Different CD spectral measurements were taken for native
DtTCI, and DtTCI exposed to (i) thermo-pH extremities and (ii)
treated with 6 M guanidine (Gdn) HCl and DTT over the range of
190–250 nm (protein solutions of 0.5 mg/ml) following Leach and
Fish (1977). The CD spectra were collected on a Jasco spectropolar-
imeter (J-810 Circular Dichroism System) equipped with a stopped
flow chamber and thermostated cell holder with 1 nm bandwidth,
a scanning speed of 50 nm/min at cell length of 0.2 cm and at 25 °C
temperature (over three accumulations) in 10 mM Tris/HCl, pH 8
(Bhattacharyya and Babu, 2006). The molar ellipticity was obtained
and the percent secondary structure was predicted using the pro-
gram k2d (Yang et al., 1986; Andrade et al., 1993).
3.13. Parasite culture
P. falciparum lines 3D7 and FCR3 cultures were maintained and
synchronized in vitro using the method of Trager and Jensen
(1976). The cultures of P. falciparum were maintained with human
O
+
RBCs suspended in 1complete culture medium [RPMI 1640
containing 0.5% AlbuMAX-I (Invitrogen) and 0.2% sodium bicar-
bonate, gentamicin (25 mg/l) and hypoxanthine 0.4 mM]. The
two culture lines of P. falciparum cultures were synchronized twice
by sorbitol treatment, and late stage trophozoites were enriched by
flotation on 65% percoll, respectively, as outlined by Kennedy et al.
(2002). The parasitemia of the mature trophozoite enriched prepa-
rations were determined by Giemsa staining and preparations of
over 90% purity were used in the erythrocyte invasion inhibition
assay.
3.14. Growth inhibition assay (GIA)
Following enumeration of parasitemia using Giemsa, the ongo-
ing culture was diluted using 2complete culture medium (dou-
ble concentration of the basic constituents of the 1culture
medium) to a final parasitemia of 0.3 ± 0.1% with a 2% hematocrit
for the assay. The assays were conducted in 96-well plates. DtTCI
was serially diluted in PBS, and 50
l
l was added to sterile flat-bot-
tomed microtitre wells (Nunc). Chymostatin, a chymotrypsin like
serine proteinase inhibitor was used as a positive control following
Dutta et al. (2003). To each well, 50
l
l of synchronized, mature tro-
phozoite-stage parasites were added (2% hematocrit, 0.3% parasite-
mia). Uninfected RBC controls (Blank) and negative controls (only
PBS, without DtTCI) were also included in the assay. The assay mix-
ture was incubated for 42 h at 37 °C in a moist atmosphere of 94%
N
2
,1%O
2
and 5% CO
2
. A 50-
l
l volume of parasites was then
washed with 250
l
l ice-cold PBS following which, 240
l
l of super-
natant was aspirated out of each well and the plate was frozen at
20 °C for at least 2 h before performing LDH assay for determina-
tion of relative parasitemia levels.
3.15. Lactate dehydrogenase (LDH) assay
Following thawing of the frozen plates, relative parasitemia lev-
els were determined by assaying for parasite lactate dehydroge-
nase activity (Basco et al., 1995; Kennedy et al., 2002). Briefly,
immediately prior to use, 100
l
l of lactate dehydrogenase assay
710 A. Bhattacharyya, C.R. Babu / Phytochemistry 70 (2009) 703–712
Author's personal copy
buffer (0.1 M Tris [pH 8.0] containing 50 mM sodium
L
-lactate,
0.25% Triton X-100, 5
l
g of 3-acetylpyridine adenine dinucleotide,
0.1 units of diaphorase and 20
l
g of Nitro Blue Tetrazolium) (all re-
agents were from Sigma) was added to each well. After addition,
the plate was instantly spun (5–15 s) at 1800gto eliminate bub-
bles. Next the plate was incubated for 30 min in the dark at room
temperature, with shaking. Absorbance was measured at 650 nm,
and percent inhibition of invasion was calculated as follows:
Assays were done in triplicate, and error bars represent the
standard deviations.
%Inhibition ¼100%
A
650
DtTCI treated sample A
650
RBC only 100
A
650
infected RBC A
650
RBC only

:
3.16. Statistical analysis
For the pLDH based assays, all values were normalized to per-
cent control activity and 50% inhibitory concentrations (IC
50
s) were
calculated by non-linear regression with the Prism 5.1 program
(GraphPad Software).
Acknowledgements
We thank the Ministry of Environment and Forests, Govt. of
India for financial support; Forest Departments at Campbell Bay
(Great Nicobar) for help at the field station; Andaman Public Works
Department for their logistic support; Dr. Dinkar Salunke, National
Institute of Immunology, New Delhi for the N-terminal sequencing
and MALDI-TOF mass spectrometry; Mr. A. Hafiz and Mr. M. Alam,
ICGEB New Delhi, for help in the parasite culture. Arindam Bhatta-
charyya is thankful to the University Grants Commission (UGC),
Govt. of India for Junior Research Fellowship.
References
Andrade, M.A., Chacón, P., Merelo, J.J., Morán, F., 1993. Evaluation of secondary
structure of proteins from UV circular dichroism using an unsupervised
learning neural network. Prot. Eng. 6, 383–390.
Araujo, A.P., Hansen, D., Vieira, D.F., Oliveira, C., Santana, L.A., Beltramini, L.M.,
Sampaio, C.A., Sampaio, M.U., Oliva, M.L., 2005. Kunitz-type Bauhinia
bauhinioides inhibitors devoid of disulfide bridges: isolation of the cDNAs,
heterologous expression and structural studies. Biol. Chem. 386, 561–568.
Basco, L.K., Marquet, F., Makler, M.M., Lebras, J., 1995. Plasmodium falciparum and
Plasmodium vivax – lactate dehydrogenase activity and its application for
in vitro drug susceptibility assay. Exp. Parasitol. 80, 260–271.
Bhattacharyya, A., Babu, C.R., 2006. Exploring the protease mediated conformational
stability in a trypsin inhibitor from Archidendron ellipticum seeds. Plant Physiol.
Biochem. 44 (11-12), 637–644.
Bhattacharyya, A., Mazumdar, S., Mazumdar-Leighton, S., Babu, C.R., 2006. A Kunitz
proteinase inhibitor from Archidendron ellipticum seeds: purification,
characterization and kinetic properties. Phytochemistry 67, 232–241.
Birk, Y., 1976. Trypsin and chymotrypsin inhibitors from soybeans. Methods
Enzymol. 45, 700–707.
Birk, Y., 2003. Inhibitor families of plant origin: classification and characterization.
In: Plant Protease Inhibitors. Springer Verlag, Berlin, pp. 11–74.
Blackman, M.J., 2000. Proteases involved in erythrocyte invasion by the malaria
parasite: function and potential as chemotherapeutic targets. Curr. Drug Targets
1, 59–83.
Bradford, M.M., 1976. A rapid and sensitive method for the quantitation of
micrograms quantities for proteins utilizing the principle of protein–dye
binding. Anal. Biochem. 72, 248–254.
Chaudhary, N.S., Shee, C., Islam, A., Ahmad, F., Yernool, D., Kumar, P., Sharma, A.K.,
2008. Purification and characterization of a trypsin inhibitor from Putranjiva
roxburghii seeds. Phytochemistry 69 (11), 2120–2126.
Crestfield, A.M., Moore, S., Stein, W.H., 1963. The preparation and enzymatic
hydrolysis of reduced and S-carboxymethylated proteins. J. Biol. Chem. 238,
622–627.
Datta, K., Usha, R., Dutta, S.K., Singh, M., 2001. A comparative study of the winged
bean protease inhibitors and their interaction with proteases. Plant Physiol.
Biochem. 39, 949–959.
de Leo, F., Volpicella, M., Liuni, S., Gallerani, R., Ceci, L.R., 2002. PLANT-PI database
for plant proteinase inhibitors. Nucleic Acids Res. 30, 347–348.
Dejkriengkraikhul, P., Wilairat, P., 1983. Requirement of malarial proteases in the
invasion of human red cells by merozoites of Plasmodium falciparum. Parasitol.
Res. 69 (3), 313–317.
Dutta, S., Haynes, J.D., Moch, J.K., Barbosa, A., Lanar, D.E., 2003. Invasion-inhibitory
antibodies inhibit proteolytic processing of apical membrane antigen 1 of
Plasmodium falciparum merozoites. Proc. Natl. Acad. Sci. 100 (21), 12295–
12300.
Dutta, S., Haynes, J.D., Barbosa, A., Ware, L.A., Snavely, J.D., Moch, J.K., Thomas, A.W.,
Lanar, D.E., 2005. Mode of action of invasion-inhibitory antibodies directed
against apical membrane antigen 1 of Plasmodium falciparum. Infect. Immun.
73, 2116–2122.
Garcia, V.A., Freire, M.G.M., Novello, J.C., Marangoni, S., Macedo, M.L.R., 2004.
Trypsin inhibitor from Poecilanthe parviflora seeds: purification,
characterization, and activity against pest proteases. Prot. J. 23, 343–350.
Gomes, C.E., Barbosa, A.E., Macedo, L.L., Pitanga, J.C., Moura, F.T., Moura, R.M.,
Oliveira, A.S., Queiroz, A.F., Macedo, F.P., Andrade, L.B., Vidal, M.S., Sales, M.P.,
2005. Effect of trypsin inhibitor from Crotalaria pallida seeds on Callosobruchus
maculatus (cowpea weevil) and Ceratitis capitata (fruit fly). Plant Physiol.
Biochem. 43, 1095–1102.
Gustafson, K.R., Sowder, R.C., Henderson, L.E., Parsons, I.C., Kashman, Y., Cardellina,
J.H., McMahon, J.B., Buckheit, R.W., Pannell, L.K., Boyd, M.R., 1994. Circulins A
and B. Novel human immunodeficiency virus (HIV)-inhibitory macrocyclic
peptides from the tropical tree Chassalia parvifolia. J. Am. Chem. Soc. 116 (20),
9337–9338.
Haq, S.K., Khan, R.H., 2003. Characterization of a proteinase inhibitor from Cajanus
cajan (L.). J. Prot. Chem. 22, 543–554.
Howell, S.A., Wells, I., Fleck, S.L., Kettleborough, C., Collins, C., Blackman, M.J., 2003.
A single malaria merozoite serine protease mediates shedding of multiple
surface proteins by juxtamembrane cleavage. J. Biol. Chem. 278, 23890–
23898.
Kato, K., Mayer, D.C., Singh, S., Reid, M., Miller, L.H., 2005. Domain III of Plasmodium
falciparum apical membrane antigen 1 binds to the erythrocyte membrane
protein Kx. Proc. Natl. Acad. Sci. (USA) 102 (15), 5552–5557.
Kennedy, A.R., 1998. The Bowman–Birk inhibitor from soybeans as an
anticarcinogenic agent. Am. J. Clin. Nut. 68, 1406S–1412S.
Kennedy, M.C., Wang, J., Zhang, Y., Miles, A.P., Chitsaz, F., Saul, A., Long, C.A., Miller,
L.H., Stowers, A.W., 2002. In vitro studies with recombinant Plasmodium
falciparum apical membrane antigen 1 (AMA1): production and activity of an
AMA1 vaccine and generation of a multiallelic response. Infec. Immun. 70 (12),
6948–6960.
Klemba, M., Goldberg, D.E., 2002. Biological roles of proteases in parasitic protozoa.
Annu. Rev. Biochem. 71, 275–305.
Kouzuma, Y., Suetake, M., Kimura, M., Yamasaki, N., 1992. Isolation and primary
structure of proteinase inhibitors from Erythrina variegata (Linn.) var. Orientalis
seeds. Biosci. Biotechnol. Biochem. 56, 1819–1824.
Laemmli, U.K., 1970. Cleavage of structural proteins during the assembly of the
head of bacteriophage T4. Nature 227, 680–685.
Leach, B.S., Fish, W.W., 1977. Resistance of soybean trypsin inhibitor (Kunitz) to
denaturation by guanidinium chloride. J. Biol. Chem. 252 (15), 5239–5243.
Lehle, K., Konert, U., Stern, A., Popp, A., Jaenicke, R., 1996. Effect of disulfide bonds on
the structure, function, and stability of the trypsin/tPA inhibitor from Erythrina
caffra: site-directed mutagenesis, expression, and physiochemical
characterization. Nat. Biotech. 14, 476–480.
Liao, H., Ren, W., Kang, Z., Jiang, J.H., Zhou, X.J., Du, L.F., 2007. A trypsin inhibitor
from Cassia obtusifolia seeds: isolation, characterization and activity on Pieris
rapae. Biotech. Lett. 4, 653–658.
Lingaraju, M.H., Gowda, L.R., 2008. A Kunitz trypsin inhibitor of Entada scandens
seeds: another member with single disulfide bridge. Biochim Biophys Acta.
1784 (5), 850–855.
Macedo, M.L.R., Freire, M.G.M., Cabrini, E.C., Toyama, M.H., Novello, J.C., Marangoni,
S., 2003. A trypsin inhibitor from Peltophorum dubium seeds active against pest
proteases and its effect on the survival of Anagasta kuehniella (Lepidoptera:
Pyralidae). Biochim. Biophys. Acta. 1621, 170–181.
Macedo, M.L., Garcia, V.A., Freire, M.G., Richardson, M., 2007. Characterization of a
Kunitz trypsin inhibitor with a single disulfide bridge from seeds of Inga laurina
(SW.) Willd. Phytochemistry 68, 1104–1111.
Norioka, N., Hara, S., Ikenaka, T., Abe, J., 1988. Distribution of the Kunitz and
Bowman-Birk family proteinase inhibitor in Leguminosae seeds. Agric. Biol.
Chem. 52, 1245–1252.
Oliveira, A.S., Migliolo, L.R., Aquino, O., Ribeiro, J.K.C., Macedo, L.L.P., Andrade, L.B.S.,
Bemquerer, M.P., Santos, E.A., Kiyota, S., Sales, M.P., 2007. Identification of a
Kunitz-type proteinase inhibitor from Pithecellobium dumosum seeds with
insecticidal properties and double activity. J. Agric. Food Chem. 55 (18), 7342–
7349.
Onesti, S., Brick, P., Blow, D.M., 1991. Crystal structure of a Kunitz-type trypsin
inhibitor from Erythrina caffra seeds. J. Mol. Biol. 217, 153–176.
Pachebat, J.A., Kadekoppala, M., Grainger, M., Dluzewski, A.R., Gunaratne, R.S., Scott-
Finnigan, T.J., Ogun, S.A., Ling, I.T., Bannister, L.H., Taylor, H.M., Mitchell, G.H.,
Holder, A.A., 2007. Extensive proteolytic processing of the malaria parasite
merozoite surface protein 7 during biosynthesis and parasite release from
erythrocytes. Mol. Biochem. Parasitol. 151 (1), 59–69.
Pattanaik, P., Jain, B., Ravindra, G., Gopi, H.N., Pal, P.P., Balaram, H., Balaram, P., 2003.
Stage-specific profiling of Plasmodium falciparum proteases using an internally
quenched multispecificity protease substrate. Biochem. Biophys. Res. Com. 309,
974–979.
A. Bhattacharyya, C.R. Babu / Phytochemistry 70 (2009) 703–712 711
Author's personal copy
Ramasarma, P.R., Rao, A.G.A., Rao, D.R., 1995. Role of disulfide linkages in structure
and activity of proteinase inhibitor from horsegram (Dolichos biflorus). Biochim.
Biophys. Acta 1248, 35–42.
Richardson, M., 1977. The proteinase inhibitors of plants and microorganisms.
Phytochemistry 16, 159–169.
Richardson, M., 1991. Seed storage proteins: the enzyme inhibitors. In: Rogers, L.J.
(Ed.), Methods in Plant Biochemistry, Amino Acids, Proteins and Nucleic Acids,
vol. 5. Academic Press, New York, pp. 59–305.
Rockett, K.A., Playfair, J.H.L., Ashall, F., Targetl, G.A.T., Anjliker, H., Shaw, E., 1990.
Inhibition of intraerythrocytic development of Plasmodium falciparum by
proteinase inhibitors. FEBS Lett. 259, 257–259.
Rosenthal, P.J., 2002. Hydrolysis of erythrocyte proteins by proteases of malaria
parasites. Curr. Opin. Hematol. 9, 140–145.
Roychaudhuri, R., Sarath, G., Zeece, M., Markwell, J., 2003. Reversible denaturation
of the soybean Kunitz trypsin inhibitor. Arch. Biochem. Biophys. 412, 20–26.
Segel, I.H., 1975. Enzyme Kinetics. John Wiley and Sons, New York.
Shewry, P.R., Lucas, J.A., 1997. Plant proteins that confer resistance to pests and
pathogens. In: J. Callow (Ed.), Adv. Bot. Res. 26, 135–192.
Souza, E.M.T., Teles, R.C.L., Siqueira, E.M.A., Freitas, S.M., 2000. Effects of denaturing
and stabilizing agents on the inhibitory activity and conformational stability of
Schizolobium parayba chymotrypsin inhibitor. J. Prot. Chem 19, 507–513.
Swartz, M.J., Mitchell, H.L., Cox, D.J., Reeck, G.R., 1977. Isolation and characterization
of trypsin inhibitor from opaque-2 corn seeds. J. Biol. Chem. 252, 8105–8107.
Sweet, R.M., Wright, H.T., Janin, J., Chothia, C.H., Blow, D.M., 1974. Crystal structure
of the complex of porcine trypsin with soybean trypsin inhibitor (Kunitz) at
26°A resolution. Biochemistry 13, 4212–4228.
Tam, J.P., Lu, Y.A., Yang, J.L., Chiu, K.W., 1999. An unusual structural motif of
antimicrobial peptides containing end-to-end macrocycle and cystine-knot
disulfides. Proc. Nat. Acad. Sci. 96, 8913–8918.
Terada, S., Fujimura, S., Kino, S., Kimoto, E., 1994. Purification and characterization
of three proteinase inhibitors from Canavalia lineata seeds. Biosci. Biotechnol.
Biochem. 58, 371–375.
Tetenbaum, J., Miller, L.M., 2001. A new spectroscopic approach to examining the
role of disulfide bonds in the structure and unfolding of soybean trypsin
inhibitor. Biochemistry 40, 12215–12219.
Trager, W., Jensen, J.B., 1976. Human malaria parasite in continuous culture. Science
193, 673–675.
Worley, K.C., Culpepper, P., Wiese, B.A., Smith, R.F., 1998. BEAUTY-X: enhanced
BLAST searches for DNA queries. Bioinformatics 14 (10), 890–891.
Wu, J., Yang, J.T., Wu, C.S., 1992. Beta-II conformation of all-beta proteins can be
distinguished from unordered form by circular dichroism. Anal. Biochem. 200,
359–364.
Yang, J.T., Wu, C.S.C., Martinez, H.M., 1986. Calculation of protein conformation
from circular dichroism. Methods Enzymol. 30, 208–269.
Yoshizaki, L., Troncoso, M.F., Lopes, J.L.S., Hellman, U., Beltramini, L.M., Wolfenstein-
Todel, C., 2007. Calliandra selloi Macbride trypsin inhibitor: isolation,
characterization, stability, spectroscopic analyses. Phytochemistry 68, 2625–
2634.
Zhou, J.Y., Liao, H., Zhang, N.H., Tang, L., Xu, Y., Chen, F., 2008. Identification of a
Kunitz inhibitor from Albizzia kalkora and its inhibitory effect against pest
midgut proteases. Biotechnol. Lett. 30 (8), 1495–1499.
712 A. Bhattacharyya, C.R. Babu / Phytochemistry 70 (2009) 703–712
... We choose to focus on plants derived protease inhibitors because due to their insecticidal property that has emerged as an interesting strategy for insect pest control (Azzouz et al., 2005). Previous research discovered that legume plants especially from tropical rain forests contributed to a major repository of PIs diversity (Arindam & Cherukuri, 2009). We attempted to simplify the methods to screen and isolate protease inhibitors from the local plant family, namely Leguminosae. ...
... Their leaves however indicated a considerably significant inhibition of trypsin. The behavior is expected as even though it was from the same plant, proteinase inhibitors are usually varied in their botanical origin and also in their effect and consequences on the storage organ, vegetative tissue, and reproductive organ within the plant (Arindam & Cherukuri, 2009). ...
Article
Search for inhibitors to insect proteases is one of many strategies to control pests. Previous work has demonstrated successful purification of effective inhibitors from plant origin. Thus, the current study attempted to purify protease inhibitors from locally available medicinal plants. The study demonstrated that the ethanolic extracts of Mimosa diplotricha leaves caused a significant 80% reduction in bovine trypsin activity. The inhibitory property of the proteinaceous nature of the extract was reconfirmed through qualitative analysis using the detection of trypsin inhibitors on the SDS-PAGE technique. The ammonium precipitated trypsin inhibitor was purified using Hi-Trap G25 and resolved into a single band with a molecular weight of approximately 20.8 kDa. By using the Dixon plot the competitive inhibitor has a Ki value of 2.16 × 10-4 mM. The purified protein inhibited the protease extract of Chrysomya megacephala at IC50 of 28 μg/mL. The results highlighted the presence of trypsin inhibitor in Mimosa diplotricha and its potential as a pest control agent.
... As regards to mechanism of action, the inhibitors have shown non-competitive type of inhibition towards bovine pancreatic trypsin. Although a few like kunitz soybean trypsin inhibitor, inhibitors from Archidendron ellipticum, Putranjiva roxburghii, Pithecellobium dumosum, Piptadenia moniliformis have shown competitive type of inhibition [49][50][51][52][53], majority of the inhibitors including those from Vicia faba, Adenanthera pavonina and Achyranthes aspera followed noncompetitive inhibition kinetics [54][55][56]. ...
... Because of their small molecular masses, acidic nature (low pI value) and existence in multiple forms, the inhibitors from Abelmoschus moschatus may be categorized as Kunitz inhibitors. The presence of multiple forms of Kunitz inhibitors has been reported in Abelmoschus esculentus [21], Adenanthera pavonina [57] and Derris trifoliata [51]. A large number of proteinase inhibitors are products of multigene families [65] and the existence of multiple forms has been attributed to expression by distinct genes and /or post- ...
... Similar results were found for inhibitors Archidrendon ellipticum (Bhattacharyya et al. 2006) and Derris foliata Lour. (Bhattacharyya and Babu 2009). The stoichiometry 1:1 inhibitor:enzyme was also reported for Schizolobium parayba (Souza et al. 1995) and Derris trifoliate Lour PIs (Bhattacharyya and Babu 2009), but was different from Psophocarpus tetragonolobus, which forms complexes in a ratio of 1:2 inhibitor:chymotrypsin (Kortt 1980). ...
... (Bhattacharyya and Babu 2009). The stoichiometry 1:1 inhibitor:enzyme was also reported for Schizolobium parayba (Souza et al. 1995) and Derris trifoliate Lour PIs (Bhattacharyya and Babu 2009), but was different from Psophocarpus tetragonolobus, which forms complexes in a ratio of 1:2 inhibitor:chymotrypsin (Kortt 1980). The value of dissociation constant (K i ) calculated for EvCI (4 × 10 -8 M) is in accordance with other inhibitors, such as Derris trifoliata Lour (1.25 × 10 -10 M) (Bhattacharyya and Babu 2009), Plathymenia foliosa (1.4 × 10 -6 M) (da Silveira et al. 2008) and Schizolobium parahyba (5.85 × 10 -8 M) (Souza et al. 1995). ...
Article
Erythrina velutina is a species of arboreal leguminous that occurs spontaneously in the northeastern states of Brazil. Leguminous seeds represent an abundant source of peptidase inhibitors, which play an important role in controlling peptidases involved in essential biological processes. The aim of this study was to purify and characterize a novel Kunitz-type peptidase inhibitor from Erythrina velutina seeds and evaluate its anti-proliferative effects against cancer cell lines. The Kunitz-type chymotrypsin inhibitor was purified from Erythrina velutina seeds (EvCI) by ammonium sulphate fractionation, trypsin– and chymotrypsin–sepharose affinity chromatographies and Resource Q anion-exchange column. The purified EvCI has a molecular mass of 18 kDa with homology to a Kunitz-type inhibitor. Inhibition assays revealed that EvCI is a competitive inhibitor of chymotrypsin (with Ki of 4 × 10–8 M), with weak inhibitory activity against human elastase and without inhibition against trypsin, elastase, bromelain or papain. In addition, the inhibitory activity of EvCI was stable over a wide range of pH and temperature. Disulfide bridges are involved in stabilization of the reactive site in EvCI, since the reduction of disulfide bridges with DTT 100 mM abolished ~ 50% of its inhibitory activity. The inhibitor exhibited selective anti-proliferative properties against HeLa cells. The incubation of EvCI with HeLa cells triggered arrest in the cell cycle, suggesting that apoptosis is the mechanism of death induced by the inhibitor. EvCI constitutes an interesting anti-carcinogenic candidate for conventional cervical cancer treatments employed currently. The EvCI cytostatic effect on Hela cells indicates a promised compound to be used as anti-carcinogenic complement for conventional cervical treatments employed currently.
... As for the B-chain, we have been able to find only three examples of KSPi with antiprotozoal activity. These are aprotinin from bovine pancreas [53], a 55-amino acid peptide ShPI-I from sea anemone [54] and a 20 kDa KSPi homologue from seeds of the tropical legume liana Derris trifoliata [55]. Their antiprotozoal activity is believed to be based on their ability to inhibit serine proteases of parasites. ...
Article
Full-text available
Protozoal infections are a world-wide problem. The toxicity and somewhat low effectiveness of the existing drugs require the search for new ways of protozoa suppression. Snake venom contains structurally diverse components manifesting antiprotozoal activity; for example, those in cobra venom are cytotoxins. In this work, we aimed to characterize a novel antiprotozoal component(s) in the Bungarus multicinctus krait venom using the ciliate Tetrahymena pyriformis as a model organism. To determine the toxicity of the substances under study, surviving ciliates were registered automatically by an original BioLaT-3.2 instrument. The krait venom was separated by three-step liquid chromatography and the toxicity of the obtained fractions against T. pyriformis was analyzed. As a result, 21 kDa protein toxic to Tetrahymena as isolated and its amino acid sequence was determined by MALDI TOF MS and high-resolution mass spectrometry. It was found that antiprotozoal activity was manifested by β-bungarotoxin (β-Bgt) differing from the known toxins by two amino acid residues. Inactivation of β-Bgt phospholipolytic activity with p-bromophenacyl bromide did not change its antiprotozoal activity. Thus, this is the first demonstration of the antiprotozoal activity of β-Bgt, which is shown to be independent of its phospholipolytic activity.
... Also, very few studies focused on separating both BBI and KI from a single or specific leguminous seed variety, possibly due to the limitations such as their close molecular mass and specificity against trypsin [39,46,47]. Further, the separation of BBI and KI during the purification process is even more complex due to the oligomeric nature of BBI [48][49][50][51]. Despite these challenges, the studies of Mohanraj et al. [39] developed a two-step method to purify BBI and KI from the seeds of R. sublobata as follows (i) ammonium sulfate fractionation of CPE followed by dialysis and chromatographic (affinity and gel filtration) techniques to obtain a trypsin specific PI pool which contained both BBI as well as KI and (ii) TCA extraction of PI pool to separate BBI and KI in their pure form. ...
Article
Bowman-Birk inhibitor (BBI ~10 kDa) and Kunitz inhibitor (KI ~20 kDa) are serine protease/proteinase inhibitor(s) [PI(s)] ubiquitously found in several Leguminous plant species with insecticidal and therapeutic properties. Due to narrow molecular mass differences, the separation of these inhibitors from a single seed variety is tedious. The present study is aimed to develop a rapid protocol (˂24 h) for purifying BBI and KI from legume seeds using mild trichloroacetic acid (TCA) extraction followed by trypsin-affinity chromatography. The mature seeds of Vigna radiata and Cajanus platycarpus are used as a model to purify BBI and KI using this protocol. The BBI and KI purified from the seeds of V. radiata are labeled as VrBBI & VrKI, and C. platycarpus are labeled as CpBBI & CpKI, respectively. These PIs are confirmed by immunodetection and MALDI-TOF studies and further characterized for their structural (CD & fluorescence spectroscopy) and functional properties (temperature & DTT stability). BBI(s) purified using the above process are effective in the management of castor semi-looper 'Achaea janata', while KI(s) are effective in the management of pod borer 'Helicoverpa armigera'. Besides, both BBI(s) and KI(s) have significant potential in controlling the growth of methicillin-sensitive 'Staphylococcus aureus', a gram-positive pathogenic bacterium.
... Numerous bioactive compounds have been identified from Derris trifoliata Lour. [33][34][35][36][37]. We previously isolated some rotenoid compounds from D. trifoliata Lour. ...
Article
Full-text available
We previously isolated derrisfolin A, a novel rotenoid derivative, from the stems of Derris trifoliata Lour. (Leguminosae). Here, we report that derrisfolin A induces the expression of endogenous regucalcin (RGN) protein in both pancreatic MIN6 β‐cells and RAW264.7 macrophages. Induction of RGN expression by derrisfolin A or retrovirus‐mediated gene transfer in MIN6 cells and RAW264.7 macrophages significantly decreased lipopolysaccharide (LPS)‐induced mRNA expression of Nos2, Il1b, and Tnf via nuclear factor‐κB activation; reduced LPS‐induced apoptosis in MIN6 cells, accompanied by decreased production of nitric oxide, interleukin‐1β, and tumor necrosis factor‐α; and attenuated generation of LPS‐induced reactive oxygen species, malondialdehyde, and 3‐nitrotyrosine in MIN6 cells. Additionally, in co‐cultures of MIN6 cells with RAW264.7 macrophages in the presence of LPS, induction of RGN expression by derrisfolin A or retrovirus‐mediated gene transfer in RAW264.7 macrophages attenuated apoptosis and oxidative/nitrosative stress in MIN6 cells. These results suggest that the induction of RGN expression in MIN6 cells was effective in suppressing LPS‐induced inflammatory cytotoxicity and that in co‐culture conditions, the induction of RGN expression in RAW264.7 macrophages blocked LPS‐induced paracrine effects of RAW264.7 macrophages on inflammatory cytotoxicity in MIN6 cells. Our findings suggest that derrisfolin A, a chemical inducer of RGN, might be useful for developing a new drug against macrophage‐associated β‐cell inflammation in type 2 diabetes.
... However, different parts of the plant are used in traditional medicine for treatment of wounds, calculus, rheumatism and dysmenorrhea and asthma [3]. This plant's extracts and metabolites have been found to possess significant antiinflammatory, antimicrobial, larvicidal, pesticidal, nitric oxide inhibitory, cytotoxic, anti-fungal, and cancer chemopreventive activities [4]- [7]. It has been found from previous phytochemical investigations on Derris trifoliata that the most characteristic compounds of this plant are flavonoids, including rotenoids [8]. ...
Article
Full-text available
The current study was conducted to verify the traditional medicinal use and to carry out the in-vitro antioxidant activity of various solvent extracts of Derris trifoliata (aerial part). The percentage yield of ethanol, ethyl acetate and n-hexane extracts were found 2.5% w/w. Freshly prepared extracts were subjected to preliminary phytochemical screening. All extracts revealed the presence of several important phytochemicals which might be responsible for its medicinal properties. In vitro Electron transfer (ET) reaction-based assays of ethanol, ethyl acetate and n-hexane extracts have been investigated using various model systems viz., DPPH, total phenolic, tannin and flavonoid content, ferric ion reducing antioxidant power (FRAP) and reducing power assay. Hydrogen atom transfer (HAT) reaction-based assays have been conducted using Nitric Oxide (NO) scavenging and hydrogen peroxide scavenging activity assay methods. Ethanol extract was found to possess highest DPPH (IC50=16.824 µg/ml), total phenolic content (44.51 GAE/g of dried plant extract), reducing power assay (0.387±0.0006), FRAF assay (IC50=133.51 µg/ml), hydrogen peroxide scavenging (IC50=144.888 µg/ml) and nitric oxide scavenging activities (IC50=152.655 µg/ml). Whereas ethyl acetate extract was found to possess the highest total tannin content (42.56 GAE/g of dried plant extract) and total flavonoid content (78.08 QE/g of dried plant extract). In vitro antioxidant study was also performed in terms of chelation power on ferrous ions. The highest chelation power was found for ethyl acetate extract (IC50=62.489 µg/ml). The above study suggests that Derris trifoliata may be a vital source of nutraceuticals.
... In addition to legumes, several Indonesian plants contain SPI compounds, including moringa (Moringa oleifera; leaves and seeds), bitter melon (Momordica charantia; seeds), cucumber (Cucumis sativus; fruit), pumpkin (Cucurbita moschata; fruit), pineapple (Ananas comosus; fruit), sweet potato (Ipomoea batatas; tuber), and potato (Solanum tuberosum; tuber) (Bijina et al., 2011;Srikanth and Chen, 2016). Furthermore, SPI has been reported in the mangrove plant Excoecaria agallocha (Li et al., 2012) and mangrove-associated Derris trifoliata (Bhattacharyya and Babu, 2009). ...
Article
Full-text available
The outbreak of coronaviruses (CoVs) presents an enormous threat to humans. To date, no new therapeutic drugs or vaccines licensed to treat human coronaviruses remain undiscovered. This mini-review briefly reports the number of potential plants widely distributed in Indonesia for further research and development as anti-SARS-CoV-2 agents and the critical targets for SARS-CoV-2 therapy, such as angiotensin-converting enzyme 2 (ACE-2) receptor, spike protein, 3-chymotrypsin-like protease (3CLpro), papain-like protease (PLpro), RNA-dependent RNA polymerase (RdRp), helicase, and serine protease. Indonesia is rich in medicinal plants (herbal); it also has a long history of using plants to treat various hereditary diseases. However, since SARS-CoV-2 is a new disease, it has no history of plant-based treatment anywhere in the world. This mini-review describes natural products from several Indonesian plants that contain compounds that could potentially prevent or reduce SARS-CoV-2 infection, act as potential targeted therapy, and provide new therapeutic strategies to develop SARS-CoV-2 countermeasures.
Article
Full-text available
Protease inhibitors (PIs) are important biotechnological tools of interest in agriculture. Usually they are the first proteins to be activated in plant-induced resistance against pathogens. Therefore, the aim of this study was to characterize a Theobroma cacao trypsin inhibitor called TcTI. The ORF has 740 bp encoding a protein with 219 amino acids, molecular weight of approximately 23 kDa. rTcTI was expressed in the soluble fraction of Escherichia coli strain Rosetta [ DE3 ] . The purified His-Tag rTcTI showed inhibitory activity against commercial porcine trypsin. The kinetic model demonstrated that rTcTI is a competitive inhibitor, with a Ki value of 4.08 × 10 –7 mol L ⁻¹ . The thermostability analysis of rTcTI showed that 100% inhibitory activity was retained up to 60 °C and that at 70–80 °C, inhibitory activity remained above 50%. Circular dichroism analysis indicated that the protein is rich in loop structures and β-conformations. Furthermore, in vivo assays against Helicoverpa armigera larvae were also performed with rTcTI in 0.1 mg mL ⁻¹ spray solutions on leaf surfaces, which reduced larval growth by 70% compared to the control treatment. Trials with cocoa plants infected with Mp showed a greater accumulation of TcTI in resistant varieties of T. cacao , so this regulation may be associated with different isoforms of TcTI. This inhibitor has biochemical characteristics suitable for biotechnological applications as well as in resistance studies of T. cacao and other crops.
Article
Full-text available
Protease inhibitors (PIs) are ubiquitous regulatory proteins present in all kingdoms. They play crucial tasks in controlling biological processes directed by proteases which, if not tightly regulated , can damage the host organism. PIs can be classified according to their targeted proteases or their mechanism of action. The functions of many PIs have now been characterized and are showing clinical relevance for the treatment of human diseases such as arthritis, hepatitis, cancer, AIDS, and cardiovascular diseases, amongst others. Other PIs have potential use in agriculture as insecticides, anti-fungal, and antibacterial agents. PIs from tick salivary glands are special due to their pharmacological properties and their high specificity, selectivity, and affinity to their target proteases at the tick-host interface. In this review, we discuss the structure and function of PIs in general and those PI superfamilies abundant in tick salivary glands to illustrate their possible practical applications. In doing so, we describe tick salivary PIs that are showing promise as drug candidates, highlighting the most promising ones tested in vivo and which are now progressing to preclinical and clinical trials.
Article
Full-text available
A protein proteinase inhibitor (PI) has been purified from pigeonpea Cajanus cajan (L.) PUSA 33 variety by acetic-acid precipitation, salt fractionation and chromatography on a DEAE-Cellulose column. The content of inhibitor was found to be 15 mg/20 g dry weight of pulse. The molecular weight of the inhibitor as determined by SDS-PAGE under reducing conditions was found to be about 14,000. It showed inhibitory activity toward proteolytic enzymes belonging to the serine pro-tease group, namely trypsin and ␣-chymotrypsin. The inhibitory activity was stable over a wide range of pH and temperatures. Estimation of sulfhydryl groups yielded one free cysteine and at least two disulfide linkages. N-terminal sequence homology suggests that it belongs to the Kunitz inhibitor family. Structural analysis by circular dichroism shows that the inhibitor possesses a largely disordered structure. KEY WORDS: Cajanus cajan; circular dichroism; fluorescence quenching; Kunitz inhibitor; N-terminal sequence; proteinase inhibitor.
Article
Plasmodium falciparum can now be maintained in continuous culture in human erythrocytes incubated at 38°C in RPMI 1640 medium with human serum under an atmosphere with 7 percent carbon dioxide and low oxygen (1 or 5 percent). The original parasite material, derived from an infected Aotus trivirgatus monkey, was diluted more than 100 million times by the addition of human erythrocytes at 3- or 4-day intervals. The parasites continued to reproduce in their normal asexual cycle of approximately 48 hours but were no longer highly synchronous. They have remained infective to Aotus.
Article
The trypsin and chymotrypsin inhibitory activities of two families of inhibitors, the Kunitz family and the Bowman–Birk family, were screened for in seeds of 34 leguminous species by gel filtration. Results were compared with the morphological classification of leguminous plants. There seemed to be a relationship between the inhibitors found and the evolution of leguminous plants. That is, the seeds of the more primitive leguminous species contained mainly the Kunitz family inhibitors and those of the more advanced ones had the Bowman–Birk family inhibitors. Results suggested that the Kunitz family inhibitors in leguminous seeds have gradually been replaced by the Bowman–Birk family inhibitors in the process of evolution.
Article
Enzymes are protein catalysts that accelerate the rates at which reactions approach equilibrium. Enzyme kinetics is the branch of biochemistry that deals with a quantitative description of this process, mainly, how experimental variables affect reaction rates. The variables that are studied include the concentrations of the enzymes, substrates (reactants), products, inhibitors, activators, the pH, temperature, and ionic strength. A complete kinetic analysis (together with complementary studies of equilibrium ligand binding, isotope exchange, covalent modification of amino acid side chains, etc.) can disclose most of the functional characteristics of a particular enzyme. These include (1) the specificities and affinities of the ligand subsites, (2) the order in which substrates bind and products leave, (3) the enzyme species that are intermediates in the overall reaction, (4) the magnitudes of component rate constants, (5) the possible identities of active-site residues, (6) the mode of action of an inhibitory drug, and (7) how the enzyme might be regulated in vivo. Enzyme kinetics combined with related approaches can show how the functional properties of a mutant or engineered enzyme compare to those of its wild-type parent. Many of the equations of enzyme kinetics are also applicable to other saturable biological processes, for example, membrane transport and receptor–ligand interactions.
Article
The trypsin and chymotrypsin inhibitory activities of two families of inhibitors, the Kunitz family and the Bowman–Birk family, were screened for in seeds of 34 leguminous species by gel filtration. Results were compared with the morphological classification of leguminous plants. There seemed to be a relationship between the inhibitors found and the evolution of leguminous plants. That is, the seeds of the more primitive leguminous species contained mainly the Kunitz family inhibitors and those of the more advanced ones had the Bowman–Birk family inhibitors. Results suggested that the Kunitz family inhibitors in leguminous seeds have gradually been replaced by the Bowman–Birk family inhibitors in the process of evolution.
Article
Three protease inhibitors, namely a chymotrypsin inhibitor (WbCI), a trypsin inhibitor (WbTI) and a chymotrypsin trypsin inhibitor (WbCTI) were purified to homogeneity from the imbibed winged bean (Psophocarpus tetragonolobus (L.) Dc) seeds by affinity chromatography on trypsin and chymotrypsin columns. The novelty of the purification lies in exploiting their differential degrees of interaction with their cognate proteases, which results in relatively higher yields for the three inhibitors as compared to the other methods used for purifying any one of them. All have an apparent molecular mass of about 20 kDa as do members of the Kunitz family of protease inhibitors. The amino-terminal sequence of WbCTI is identical to the sequence of the WTI-1 reported earlier by Yamamoto et al. whereas WbTI is a new protein whose N-terminal sequence has no homology with the N-terminal sequence of any known proteins. A comparative study of all three inhibitors shows the presence of WbTI and WbCTI only in the seeds, whereas the presence of WbCI in other tissues has already been documented. Synthesis and degradation during germination follow a similar pattern for both the inhibitors. The binding studies with cognate proteases show that WbCTI inhibits trypsin more strongly than chymotrypsin and it never forms a ternary complex even though it binds both proteases. On the other hand, WbTI inhibited only trypsin, in spite of its binding to a chymotrypsin-Sepharose column. The study shows that these protease inhibitors are good candidates for studying the protein-protein interaction at the molecular level.
Article
Throughout their lives, from dormant seed to maturity, plants are exposed to attack by a wide variety of potentially pathogenic microorganisms, predatory insects and other invertebrate pests. Plant resistance is often divided into constitutive defense, expressed as a normal feature of plant development, or inducible defense, switched on following contact with the invading organism. A fundamental distinction is often drawn between pre–existing or constitutive features of the plant, and inducible systems switched on following challenge by a pest or pathogen. Lectins bind to chitin and other glycoconjugates containing N–acetylglycosamine or N–acetylneuraminic acid, and also cause the agglutination of mammalian red blood cells. The most widely studied graminaceous lectin is wheat germ agglutinin, an Mr 36000 protein consisting of two identical 171–residue chains. Thionins were first purified from wheat flour in the early 1940s and called purothionins. Lysozymes from egg white and bacteriophage are among the most widely studied proteins; they inhibit bacterial growth by hydrolysis of cell–wall peptidoglycans. In devising strategies for the manipulation of defense proteins in transgenic plants it is essential to consider the behavior of the target organism, as well as the construct to be used. The specificity of protective proteins for different groups of insects and other invertebrates provides an opportunity to develop resistances that are either highly specific or have wide spectra, based on individual proteins or combinations. A combination of protective proteins may also assist in preventing the development of resistance to individual components.
Article
Recent (post-1972) advances in our knowledge of the proteinase inhibitors of plants and micro-organisms are reviewed. Details of the specificity, occurrence and distribution of these proteins are summarized, and modern methods for their isolation, purification and assay are discussed. Certain homologies revealed by comparison of the amino acid sequences of several inhibitors are noted. Details of their reactive (inhibitory) sites are tabulated and discussed in relation to the proposed mechanisms of action of these proteins. Recent experiments on the intracellular localization of the inhibitors, their physiology and possible functional roles are described. The nutritional significance, possible therapeutic use and value of the proteinase inhibitors as laboratory tools are also discussed.
Article
The CD spectrum of certain all-β globular proteins resembles that of unfolded proteins with a characteristic negative band around 200 nm. The conformation of this class is tentatively termed β-II, which had two features that were absent for unfolded proteins. First, β-II proteins usually had CD bands due to aromatic side groups in the near-ultraviolet region. Second, the CD intensities both in the far- and in the near-uv region of these compact and rigid proteins usually showed a sharp transition upon thermal denaturation, whereas those of an unordered form changed linearly with rising temperature.