ArticlePDF Available

Horse chestnut (Aesculus hippocastanum L.) starch: Basic physico-chemical characteristics and use as thermoplastic material

Authors:
Carbohydrate
Polymers
112
(2014)
677–685
Contents
lists
available
at
ScienceDirect
Carbohydrate
Polymers
j
ourna
l
ho
me
page:
www.elsevier.com/locate/carbpol
Horse
chestnut
(Aesculus
hippocastanum
L.)
starch:
Basic
physico-chemical
characteristics
and
use
as
thermoplastic
material
J.
Casta˜
noa,
S.
Rodríguez-Llamazaresb,,
K.
Contrerasb,
C.
Carrascoc,∗∗,
C.
Pozob,
R.
Bouzad,
C.M.L.
Francoe,
D.
Giraldof
aUnidad
de
Desarrollo
Tecnológico,
Universidad
de
Concepción,
Avda.
Cordillera
2634,
Coronel,
Chile
bCentro
de
Investigación
de
Polímeros
Avanzados
(CIPA),
Beltran
Mathiew
224,
Concepción,
Chile
cMaterials
Engineering
Department,
Universidad
de
Concepción,
Edmundo
Larenas
270,
Concepción,
Chile
dDepartamento
de
Física,
E.U.P.
Ferrol,
Universidad
de
A
Coru˜
na,
Avda.
19
de
Febrero,
s/n,
15405
Ferrol,
Spain
eDepartamento
de
Engenharia
e
Tecnologia
de
Alimentos,
Universidade
Estadual
Paulista,
R.
Cristóvão
Colombo,
2265
São
José
do
Rio
Preto,
SP,
Brazil
fDepartment
of
Metallurgical
and
Materials
Engineering,
Faculty
of
Engineering,
University
of
Antioquia
UdeA,
Calle
70
No.
52-21,
Medellín,
Colombia
a
r
t
i
c
l
e
i
n
f
o
Article
history:
Received
29
March
2014
Received
in
revised
form
31
May
2014
Accepted
8
June
2014
Available
online
28
June
2014
Keywords:
Aesculus
hippocastanum
L.
Amylopectin
structure
Thermoplastic
starch
Non-edible
source
a
b
s
t
r
a
c
t
Starch
isolated
from
non-edible
Aesculus
hippocastanum
seeds
was
characterized
and
used
for
prepar-
ing
starch-based
materials.
The
apparent
amylose
content
of
the
isolated
starch
was
33.1%.
The
size
of
starch
granules
ranged
from
0.7
to
35
m,
and
correlated
with
the
shape
of
granules
(spherical,
oval
and
irregular).
The
chain
length
distribution
profile
of
amylopectin
showed
two
peaks,
at
polymeriza-
tion
degree
(DP)
of
12
and
41–43.
Around
53%
of
branch
unit
chains
had
DP
in
the
range
of
11–20.
A.
hippocastanum
starch
displayed
a
typical
C-type
pattern
and
the
maximum
decomposition
temperature
was
317 C.
Thermoplastic
starch
(TPS)
prepared
from
A.
hippocastanum
with
glycerol
and
processed
by
melt
blending
exhibited
adequate
mechanical
and
thermal
properties.
In
contrast,
plasticized
TPS
with
glyc-
erol:malic
acid
(1:1)
showed
lower
thermal
stability
and
a
pasty
and
sticky
behavior,
indicating
that
malic
acid
accelerates
degradation
of
starch
during
processing.
©
2014
Elsevier
Ltd.
All
rights
reserved.
1.
Introduction
The
growing
accumulation
of
plastic
residues
in
landfills
and
oceans
and
the
strong
dependency
of
plastic
production
on
fossil
petroleum
resources
have
stimulated
research
on
bio-based
plas-
tic
alternatives.
Starch
is
a
renewable
and
inherently
biodegradable
raw
material
of
low-cost
that
can
be
transformed
into
bio-plastics
with
technologies
currently
used
for
manufacturing
synthetic
plastics.
The
worldwide
production
of
starch-based
plastics
is
esti-
mated
to
exceed
150,000
tons/year
(European
Bioplastic,
2014).1
Thermoplastic
starch
(TPS)
can
be
obtained
by
physical
mod-
ification
of
starch,
mixing
starch
with
a
plasticizer,
with
and
without
shear
stress,
at
temperatures
over
100 C.
Water,
the
main
Corresponding
author.
Tel.:
+56
41
2661216;
fax:
+56
41
2751233.
∗∗ Corresponding
author.
Tel.:
+56
41
2207170;
fax:
+56
41
2203391.
E-mail
addresses:
s.rodriguez@cipachile.cl
(S.
Rodríguez-Llamazares),
ccarrascoc@udec.cl
(C.
Carrasco).
1Bioplastic
facts
and
figures.
European
Bioplastic
2013
(http://en.european-
bioplastics.org/multimedia/).
plasticizer
of
TPS,
is
used
together
with
other
plasticizers,
such
as
polyols
(sorbitol,
xylitol,
pentaerythritol
and
glycerol),
sugar
alco-
hols
and
non-ionic
and
anionic
surfactants
(Vieira,
da
Silva,
dos
Santos,
&
Beppu,
2011).
There
are
many
reports
of
starch
coming
from
different
botanical
sources
to
prepare
TPS.
In
most
studies
starch
comes
from
edible
fruits,
vegetables,
seeds
and
so
on
(van
Soest,
Benes,
de
Wit,
&
Vliegenthart,
1996;
Zullo
&
Iannace,
2009).
The
use
of
edible
sources
for
fabrication
of
plastics
is
controversial,
especially
in
developing
countries
with
a
deficit
of
arable
land
and
food.
Therefore,
non-edible
starch
has
gained
much
attraction
for
use
in
starch-based
materials
(Bharadwaj,
2010).
TPS
has
limited
use
in
the
plastic
industry
because
its
mechani-
cal
properties
depends
on
the
humidity
condition,
(Bertuzzi,
Castro
Vidaurre,
Armada,
&
Gottifredi,
2007;
Han,
2014,
chap.
9).
An
alternative
to
reduce
water
absorption
of
TPS
is
the
chemical
modification
of
starch
before
or
during
melt
blending.
When
the
modification
is
made
during
the
extrusion,
the
process
is
known
as
reactive
extrusion.
Poly(carboxylic
acid)s
such
as
citric,
tartaric
and
malic
acids
have
been
used
as
cross-linking
agents
and
co-
plasticizers
to
elaborate
TPS.
Poly(carboxylic
acid)s
delay
water
uptake,
retrogradation
and
improve
mechanical
properties
of
TPS
http://dx.doi.org/10.1016/j.carbpol.2014.06.046
0144-8617/©
2014
Elsevier
Ltd.
All
rights
reserved.
678
J.
Casta˜
no
et
al.
/
Carbohydrate
Polymers
112
(2014)
677–685
(Shi
et
al.,
2007)
and
are
also
non-toxic
and
of
low-cost.
Zuraida,
Yusliza,
Anuar,
and
Mohd
Khairul
Muhaimin
(2012)
reported
the
preparation
of
TPS
from
Malaysian
sago
and
different
ratios
of
citric
acid
and
glycerol.
TPS
prepared
with
glycerol:citric
acid
ratio
of
1:1
showed
higher
tensile
strength
than
those
prepared
only
with
glycerol.
Jiugao,
Ning,
and
Xiaofei
(2005)
showed
by
FTIR
spec-
troscopy
and
X-ray
diffractometry
that
citric
acid
reduced
hydrogen
bonding
between
starch
and
glycerol
during
storage,
inhibiting
the
formation
of
amylose–glycerol
inclusion
complexes
and
therefore
the
formation
of
V-type
structures
associated
with
re-crystallized
starch
(i.e.
retrogradation).
Citric,
malic
and
tartaric
acids
were
also
used
as
a
cross-linking
agent
of
starch/poly(butyene
adipate
co-terephthalate)
blends
to
improve
the
compatibility
of
the
poly-
meric
phase.
However,
these
organic
acids
promote
the
hydrolysis
of
starch,
degrading
them
to
dextrin
and
fragments
of
low
molecu-
lar
weight
(Olivato,
Grossmann,
Yamashita,
Eiras,
&
Pessan,
2012).
It
would
be
interesting
to
explore
the
effect
of
malic
acid
on
TPS
plasticized
by
glycerol,
because
of
its
lower
tendency
to
pro-
mote
chemical
modification
of
starch
(Olivato,
Grossmann,
Bilck,
&
Yamashita,
2012;
Olivato
et
al.,
2012b;
Ruxanda
&
Teac ´
ˆ
a,
2012).
An
attractive
source
of
non-edible
starch
is
the
seed
of
horse
chestnut
(Aesculus
hippocastanum
L.),
which
a
starch
content
over
35
wt.%
(ˇ
Cukanovi´
c
et
al.,
2011).
The
horse
chestnut
was
introduced
in
Chile
around
XVIII
century
with
arrival
of
European
immigrants,
and
due
to
its
exceptional
adaptability
has
spread
throughout
the
whole
country.
It
is
often
found
as
ornamental
trees
in
urban
communities
and
in
grassland
of
Chilean
fields.
Currently,
seeds
of
A.
hippocastanum
are
used
for
ornamental
purposes,
but
most
of
them
end
up
in
landfills.
Most
reports
related
to
seeds
of
A.
hippocastanum
are
focused
on
their
biochemical
composition
and
morphoanatomical
character;
however,
we
have
not
found
detailed
studies
about
the
chemical
and
physical
characteristics
of
A.
hip-
pocastanum
starch.
In
this
work,
A.
hippocastanum
starch
was
characterized
and
used
to
prepare
TPS
by
the
melt
blending
method.
We
studied
the
rheological,
mechanical
and
thermal
properties
of
this
TPS
and
the
effect
of
incorporating
malic
acid
as
co-plasticizer.
In
addi-
tion,
we
compared
the
morphological
and
mechanical
properties
of
TPS
from
A.
hippocastanum
starch
with
TPS
from
starch
of
Arau-
caria
araucana
seeds,
an
edible
source
with
C-type
semi-crystalline
structure
and
excellent
precursor
for
starch-based
materials
with
adequate
performance
(Casta˜
no,
Bouza,
Rodríguez-Llamazares,
Carrasco,
&
Vinicius,
2012).
2.
Experimental
2.1.
Materials
Starch
of
ripe
A.
hippocastanum
seeds,
collected
at
the
campus
of
the
University
of
Concepción,
Chile,
was
isolated
by
the
wet
milling
method.
The
plasticizers
used
were
glycerol
(99.5%
purity)
and
dl-
malic
acid
(99%
purity)
supplied
by
Oceanquímica
S.A.
(Valparaíso,
Chile)
and
Sigma
Aldrich
(Santiago,
Chile),
respectively.
2.2.
Characterization
of
A.
hippocastanum
starch
Physical–chemical
and
morphological
features
of
A.
hippocas-
tanum
starch
granules
are
summarized
in
Table
1.
Ash
and
moisture
content
were
determined
gravimetrically
in
triplicate
according
to
ASTM
D1102-56
and
Chilean
Standard
176/1
of
84,
respectively.
Nitrogen
content
was
assessed
by
using
Kjendhal
technique
(AOAC,
1990).
Swelling
power
(g/g
of
starch
on
dry
weight)
and
solubility
(g/g
of
starch
on
dry
weight)
were
determined
in
triplicate
using
the
method
reported
by
Singh,
McCarthy,
Singh,
Moughan,
and
Kaur
(2007).
In
short,
a
suspension
of
A.
hippocastanum
starch
was
Table
1
Physico-chemical
and
morphological
features
of
A.
hippocastanum
starch.
%
Small
granule
size
(0.5–10
m)a30.7
%
Medium
granule
size
(11–30
m)a67.6
%
Large
granule
size
(>31
m)a1.7
Moisture
contentb7.1
±
0.1
Ash
content
0.20
Apparent
amylose
content
(%)
33
Nitrogen
content
0.5
±
0.1
Phosphorus
content
(%)
0.0118
Calcium
content
(%) 0.0179
Sulfur
content
(%)
0.0033
Potassium
content
(%)
0.0465
Swelling
power
(g/g
of
starch
on
dry
weight)
6.5
±
0.2
Solubility
(g/g
of
starch
on
dry
weight)
3.1
±
0.1
Color
White
aDetermined
by
laser
diffraction
particle
size
analyzer.
bDeterminated
by
thermogravimetric
analyzer.
heated
at
60 C
for
30
min
followed
by
centrifugation
at
4000
rpm
for
30
min.
The
supernatant
was
decanted
and
dried
over
night
at
70 C
(soluble
solid).
The
solid
residue
of
centrifugation
was
weighed
for
swelling
power
measurement.
The
apparent
amylose
content
was
assessed
by
using
a
col-
orimetric
method
according
to
EN
ISO
6647
parts
I
and
II,
which
is
based
on
amylose–iodine
complex
formation.
The
absorbance
was
measured
at
620
nm
by
using
an
UV
spectrophotometer
(UV
mini
1240
Shimadzu,
Japan).
The
particle
size
distribution
was
determined
with
a
laser
diffraction
particle
size
analyzer
(Beck-
man
Coulter
LSTM 200,
Brea
CA,
USA).
At
least
three
measurements
were
done
to
obtain
the
average
particle
size
distribution
curves.
The
phosphorus,
sulfur,
calcium
and
potassium
content
were
quan-
tified
by
X-ray
fluorescence
spectrometry
(S4
Pioneer,
Bruker-AXS,
Germany).
5
g
of
powdered
starch
were
mounted
on
prolene
films,
and
analyzed
under
helium
atmosphere.
The
results
were
evaluated
by
SpectraPlus
v.1.6
spectrometer
software
(Socabim).
Thermal
stability
of
A.
hippocastanum
starch
was
evaluated
by
thermo-
gravimetric
analyses
(TGA).
Morphological
of
starch
granules
was
examined
with
scanning
electron
microscopy.
X-ray
diffraction
(XRD)
was
used
to
analyze
crystalline
structure
of
starch
granules.
A.
hippocastanum
starch
was
debranched
using
isoamylase
3U
(Megazyme
International,
Ireland)
enzyme
according
to
Wong
and
Jane
(1995).
The
branch
chain
length
distribution
of
amy-
lopectin
was
analyzed
according
to
Moraes,
Alves,
and
Franco
(2013)
using
high-performance
anion-exchange
chromatography
with
pulsed
amperometric
detector
(HPAEC-PAD)
system
(ICS
3000,
Dionex
Corporation,
Sunnyvale,
USA)
equipped
with
an
AS40
automatic
sampler.
Samples
were
filtered
(0.22
m
membrane)
and
injected
into
the
HPAEC-PAD
system
(20
L
sample
loop).
The
flow
rate
was
0.8
mL/min
at
40 C.
The
standard
quadru-
ple
potential
(E)
waveform
was
employed
with
the
following
periods
and
pulse
potentials:
E1
=
0.10
V
(t1=
0.40
s);
E2
=
2.00
V
(t2=
0.02
s);
E3
=
0.60
V
(t3=
0.01
s);
E4
=
0.10
V
(t4=
0.06
s).
All
elu-
ents
were
prepared
with
ultrapure
water
(18
m
cm)
with
N2
sparging.
Eluent
A
was
150
mM
NaOH
and
eluent
B
was
500
mM
sodium
acetate
and
150
mM
NaOH.
The
branched
chains
of
amy-
lopectin
were
separated
using
a
Dionex
CarboPacTM
PA-100
guard
column
(4
×
50
mm)
and
a
Dionex
CarboPacTM
PA-100
column
(4
×
250
mm).
The
gradient
of
eluent
B
was
28%
at
0
min,
40%
at
15
min,
and
72%
at
105
min.
The
data
were
analyzed
using
the
Chromeleon
software,
version
6.8
(Dionex
Corporation,
USA).
The
analyses
were
performed
in
duplicate.
2.3.
Preparation
of
thermoplastic
starch
Before
melt-blending,
A.
hippocastanum
starch
and
plasticizers
were
premixed
by
hand
at
room
temperature.
A
Brabender
internal
mixer
(Plastograph®EC
plus,
Mixer
50EHT32,
Germany)
was
used
J.
Casta˜
no
et
al.
/
Carbohydrate
Polymers
112
(2014)
677–685
679
Table
2
Composition,
rheological
and
tensile
properties
of
TPS.
Sample
code
A.
hippocas-
tanum
starch
(wt.%)
A.
araucana
starch
(wt.%)
Malic
acid
(wt.%)
Glycerol
(wt.%)
Torque
maximum
(N
m)
Plasticization
energy
(N
m/min)
Steady
state
(N
m)
Tensile
modulus
(MPa)
Tensile
strength
(MPa)
Elongation
at
break
(%)
CG-30
70
30
30.5
9.0
15.1
3.0
±
0.3a1.5
±
0.1
1.02
±
0.04
1.5
±
0.2b
0.3
±
0.2c
CG-MA-30
70
15
15
38.5
14.6
6.8
PG-30
70
30
22.8
8.8
17.4
4.0
±
0.5a2.4
±
0.2
1.14
±
0.07
2.0
±
0.3b
0.8
±
0.2c
PG-MA-30
70
15
15
31.5
10.9
4.4
a10%.
b30%.
c90%
at
unitary
deformation
of
strain.
to
prepare
TPS.
The
samples
were
blended
at
120 C
and
60
rpm
for
15
min
and
torque
variation
as
a
function
of
time
was
recorded.
The
torque
values
of
maximum
peak
and
steady
state,
and
the
plasti-
cization
energy
were
determined
from
the
torque
rheometer
data
(see
Table
2).
The
integral
of
the
torque
curve,
from
time
zero
(when
all
the
components
were
in
the
mixing
chamber)
to
the
time
where
the
maximum
torque
is
reached,
was
reported
as
the
plasticization
energy
(Casta˜
no,
Rodríguez-Llamazares,
Carrasco,
&
Bouza,
2012).
Samples
were
labeled
specifying
the
type
of
starch
and
the
type
and
wt.%
of
the
plasticizer
agent.
For
example,
in
the
label
CG-MA-
30
C
corresponded
to
A.
hippocastanum
starch,
G
to
glycerol,
MA
to
malic
acid
and
30
to
the
wt.%
of
plasticizers
used
to
prepare
TPS,
i.e.
in
this
case
15
wt.%
of
glycerol
and
15
wt.%
of
malic
acid.
Table
2
summarizes
composition
of
thermoplastic
starch.
CG-30
and
PG-30
samples
for
tension
tests
were
injection-
molded
by
using
a
Minijet
II
(Haake
Thermo
Scientific,
Waltham,
USA),
under
the
following
conditions:
130 C
of
cylinder
tempera-
ture,
55 C
of
mold
temperature,
and
300
bar
of
injection
pressure.
To
compare
mechanical
and
morphological
properties
of
TPS
from
A.
hippocastanum
and
from
A.
araucana,
the
latter
was
prepared
using
the
same
condition
as
described
above.
2.4.
Characterization
of
thermoplastic
starch
2.4.1.
Scanning
electron
microscopy
(SEM)
Morphology
of
A.
hippocastanum
starch
granules
and
of
cryo-
genically
fractured
surface
of
the
mechanical
test
specimens
of
TPS
obtained
by
injection
molding
(sputtered
with
a
gold
coating
of
ca.
50
nm)
were
observed
at
200×,
600×
and
1000×
magnifications
with
a
field
emission
scanning
electron
microscope
(JEOL
JSM
6380
LV,
Tokyo,
Japan)
operated
at
15
kV.
2.4.2.
Thermogravimetric
analysis
(TGA)
Thermal
stability
of
A.
hippocastanum
starch
and
TPS
was
eval-
uated
using
a
thermogravimetric
analyzer
NETZSCH
209
F3
TGA
(Tarsus,
Selb,
Germany).
TGA
tests
were
carried
out
at
10 C/min
under
nitrogen
atmosphere
(20
mL/min),
from
30
to
900 C
for
A.
hippocastanum
starch
and
from
30
to
600 C
for
TPS.
2.4.3.
X-ray
diffraction
(XRD)
X-ray
diffraction
analysis
of
A.
hippocastanum
starch
and
TPS
was
performed
using
a
Bruker
Endeavour
diffractometer
(model
D4/max-B,
Germany)
with
Cu
K
radiation
(40
kV
and
20
mA).
The
spectra
were
recorded
over
a
2
range
of
4–35at
steps
of
0.02
and
time
per
step
of
0.02/s.
Peak
fitting
analysis
was
done
using
a
Gaussian
distribution
function.
Degree
of
crystallinity
was
cal-
culated
using
the
method
reported
by
Frost,
Kaminski,
Kirwan,
Lascaris,
and
Shanks
(2009).
2.4.4.
Fourier
transform
infrared
spectroscopy
(FTIR)
The
FTIR
spectra
of
A.
hippocastanum
starch
and
TPS
were
recorded
in
triplicate
at
room
temperature
with
a
Jasco
FTIR
4100
spectrophotometer
(USA)
in
the
range
of
4000–600
cm1at
2
cm1
resolution.
The
background
and
sample
spectra
were
scanned
64
times
in
transmittance
mode.
About
5
mg
of
finely
ground
solid
sample
was
ground
with
90
mg
of
dry
potassium
bromide
and
a
7
mm
pellet
was
formed
under
high
pressure.
Data
was
analyzed
with
spectral
manager
Version
2.
The
data
preprocessing
consisted
of
smoothing
and
manual
baseline
correction.
For
comparison,
the
spectra
were
normalized
by
using
the
band
at
1459
cm1,
related
to
the
symmetric
deformation
of
CH2of
starch
(Cael,
Koenig,
&
Blackwell,
1975),
which
does
not
change
with
treatment
(Morán,
Vázquez,
&
Cyras,
2013).
2.4.5.
Tensile
properties
Tensile
tests
were
performed
according
to
ASTM
D638
standard
on
a
Karg
Industrietechnik
universal
testing
machine
(Krailling,
Germany)
using
a
type
V
dumbbell
specimen.
The
crosshead
speed
was
set
at
5
mm/min.
Tensile
strength,
percent
of
elongation
at
break
and
Young’s
modulus
at
10%,
30%
and
90%
of
strain
were
calculated
from
the
stress–strain
curves.
At
least
7
individual
mea-
surements
were
carried
out
for
each
formulation.
3.
Results
and
discussion
3.1.
Characterization
of
A.
hippocastanum
starch
Starch
granules
of
A.
hippocastanum
seed
accumulate
funda-
mentally
in
parenchyma
cells
of
cotyledons.
The
size
of
oval
parenchyma
cells
found
was
around
86
m,
considering
only
the
large
part
of
the
cells
(see
complementary
information,
Fig.
A1).
This
value
is
similar
to
the
value
reported
by
Musatenko,
Generalova,
Martyn,
Vedenicheva,
and
Vasyuk
(2003)
for
parenchyma
cells
of
mature
A.
hippocastanum
seeds
collected
in
Kiev,
Ukraine.
Inside
cells,
a
fiber
network
associated
to
cytoskeletal
elements
was
also
observed.
Fig.
1
shows
a
SEM
image
and
the
size
distribution
of
starch
gran-
ules
isolated
from
A.
hippocastanum.
Starch
granules
in
parenchyma
cells
are
of
diverse
shape
and
size.
The
size
of
granules
is
linked
to
their
shape,
thus,
small
granules
(0.7–9.5
m)
are
spherical,
the
medium
(10–17
m)
are
oval
and
the
larger
size
(18–35
m)
are
irregular
(for
620
granules
counted).
The
granule
surface
is
smooth
without
pores
and
erosion.
A
typical
bimodal
size
distribution
of
A.
hippocastanum
starch
granules
is
shown
in
Fig.
1B.
Around
10%
of
granules
have
sizes
between
0.4
and
3.5
m
(first
interval)
and
the
remainder
granules
ranged
from
3.6
to
37
m
(second
interval).
Only
5%
of
granules
are
over
30
m
in
size.
680
J.
Casta˜
no
et
al.
/
Carbohydrate
Polymers
112
(2014)
677–685
Fig.
1.
(A)
SEM
micrograph
and
(B)
particle-size
distribution
of
A.
hippocastanum
starch.
The
apparent
amylose
content
of
isolated
A.
hippocastanum
starch
was
(33.1
±
0.1%),
considerably
higher
than
the
14.2%
and
13.8%
of
coarse
granules
and
fine
granules,
respectively,
of
A.
hippocastanum
starches
from
Bratislava,
Slovak
reported
by
Hricovíniová
and
Babor
(1992).
The
differences
in
amylose
content
can
be
attributed
to
their
different
source
and
cultivar
zone,
the
starch
isolation
procedure,
and
the
analytical
methods
used.
They
used
dimethyl
sulfoxide
in
the
isolation
procedure
and
the
amylose
content
was
determined
biamperometrically
by
sorption
of
iodine.
In
Fig.
2
a
normalized
HPAEC-PAD
chromatogram
shows
the
branch
chain
length
distribution
of
amylopectin
from
A.
hippocas-
tanum
starch.
There
are
two
peaks,
the
first
one
at
DP
=
12
(the
most
abundant
branch
unit
chain)
and
the
second
one
at
DP
41–43.
A
shoulder
at
DP
18–22
is
also
observed.
Average
DP
of
branch
chains
of
amylopectin
was
18.89
±
0.05.
The
proportion
of
amylopectin
unit
chains
of
DP
6–12,
13–24,
25–36
and
37
were
31.51%,
48.08%,
10.85%
and
9.56%,
respectively.
Previous
work
showed
correlations
among
chain
length
distribution
of
amylopectin
and
crystalline
structure
of
starch,
granule
size,
swelling
power
or
gelatinization
temperature.
Singh
et
al.
(2012)
found
that
low
size
granules
of
10–30
m
from
different
rice
bean
lines
contained
higher
amount
of
DP
11–20
branch
unit
chains
of
amylopectin
than
large
size
granules
of
30–105
m.
Around
53%
of
branch
unit
chains
of
A.
hip-
pocastanum
starch
showed
a
DP
of
11–20,
characteristic
of
small
size
granules.
The
type
of
semi-crystalline
structure
of
starch
granules
depends
on
the
DP
of
external
branch
chains
of
amylopectin
(Jane,
2009,
chap.
6).
A-type
starches
have
a
larger
proportion
of
short
B1-
and
A-chains,
which
are
located,
with
their
ends
free
and
Fig.
2.
Profile
of
amylopectin
branch
chain
length
distribution
of
A.
hippocastanum
starch
analyzed
by
HPAEC-PAD.
unrestricted,
in
one
cluster.
B-type
starches
have
longer
B-chains
(B2-,
B3-
and
B4-chains)
that
may
extend
through
multiple
clusters
and
are
restricted
in
their
movements.
Jane
et
al.
(1999)
analyzed
branch
chain-length
of
amylopectin
of
starch
isolated
from
differ-
ent
botanical
sources
and
found
some
trends
of
branch
chain-length
distribution
for
starches
with
the
same
crystalline
type.
A-type
starches
had
15.6–27.4%
of
short
chains
(DP
6–12)
and
6.6–26.7%
of
long
chains
(DP
>
37),
in
contrast
to
B-type
starches
that
had
8.5–12.3%
and
26.1–29.5%
of
short
and
long
chains,
respectively.
Similar
results
were
reported
by
Srichuwong,
Sunarti,
Mishima,
Isono,
and
Hisamatsu
(2005a),
but
with
26.4–36.3%
(narrower
range)
of
branch
chains
with
DP
9–12
for
A-type
starches.
A.
hip-
pocastanum
starch
displayed
substantial
amounts
of
short
chains
of
DP
6–12
(31.5%)
and
very
few
long
chains
(9.6%),
showing
a
diffrac-
tion
pattern
characteristic
of
semi-crystalline
structure
of
type
C,
as
will
be
discussed
later.
Each
starch
displays
its
own
chain-length
distribution
profile
of
amylopectin,
and
therefore
it
is
not
surpris-
ing
that
A.
hippocastanum
starch
is
not
following
the
trend
found
by
Jane
et
al.
(1999).
The
proportion
of
each
branch
chain
and
aver-
age
chain
length
of
A.
hippocastanum
starch
was
within
the
range
reported
for
other
C-type
starches
(Chung
&
Liu,
2012).
The
values
of
swelling
power
and
solubility
of
A.
hippocastanum
starch
are
shown
in
Table
1.
The
ability
of
starch
to
swell
in
excess
water,
as
well
as
its
solubility,
is
mainly
related
to
variations
in
chain
branch
length
distribution
of
amylopectin
and
phospho-
rus
content.
Srichuwong,
Sunarti,
Mishima,
Isono,
and
Hisamatsu
(2005b)
proposed
that
swelling
power
is
associated
with
the
unit-
chain
ratio
(APC)
of
amylopectin,
i.e.
the
ratio
of
relative
molar
distribution
of
amylopectin
unit-chains
with
DP
6–12
to
that
of
DP
6–24.
A
larger
proportion
of
short
chains
increased
the
swelling
power
of
starch
granules,
due
to
short
chains
forming
short
and
weak
double
helices,
decreasing
the
stability
of
crystalline
packag-
ing.
The
APC
ratio
of
A.
hippocastanum
starch
was
0.396,
similar
to
starch
granules
of
lesser
yam,
which
had
C-type
structure
and
its
swelling
power
at
70 C
was
around
8
g/g.
The
difference
in
swelling
properties
among
starches
is
also
attributed
to
phosphorus
compounds,
such
as
phospholipids
and
phosphate
monoesters.
Phospholipids
form
water
insoluble
com-
plexes
with
amylose
during
heating,
retarding
starch
granule
swelling,
while
phosphate
monoesters
on
long-branch
chains
of
amylopectin
decrease
the
interaction
between
double
helices,
facil-
itating
water
penetration
and
therefore
granular
swelling.
The
J.
Casta˜
no
et
al.
/
Carbohydrate
Polymers
112
(2014)
677–685
681
Fig.
3.
X-ray
diffraction
pattern
of
A.
hippocastanum
starch.
average
total
phosphorus
content
of
starch
from
A.
hippocastanum
was
0.0118%.
This
value
was
higher
than
that
reported
for
starch
from
Aesculus
assamica
(0.0016%)
cultivated
in
India
(Soni,
Sharma,
&
Bisen,
1989),
both
belonging
to
the
same
genus.
It
is
interesting
to
observe
that
the
swelling
power
of
A.
hippocastanum
starch
is
twice
of
that
of
A.
assamica
starch,
which
tentatively
can
be
attributed
to
the
lower
P
content
of
the
latter.
Unfortunately,
we
found
no
reports
about
branch
chain
length
distribution
of
amylopectin
iso-
lated
from
A.
assamica.
The
X-ray
diffraction
pattern
of
A.
hippocastanum
starch
is
shown
in
Fig.
2.
A.
hippocastanum
starch
displays
a
characteris-
tic
C-type
pattern
with
diffraction
peaks
at
2
=
5.7,
15.4,
17.2,
22.0and
22.7.
The
C-type
pattern
is
not
characteristic
of
all
starches
from
the
seed
of
genus
Aesculus.
For
example,
starch
from
Aesculus
indica
shows
an
A-type
pattern
(Wani
et
al.,
2014).
The
degree
of
crystallinity
of
A.
hippocastanum
starch
was
calculated
using
a
simple,
quick
and
objective
method
(Frost
et
al.,
2009).
Amorphous
background
scattering
was
estimated
using
an
iterative
smoothing
algorithm.
The
value
of
degree
of
crystallinity
found
was
10.7%,
which
is
lower
than
that
of
starch
from
A.
Araucana
(15.1%)
(Casta˜
noet
al.,
2012a).
Both
starches
have
a
C-type
structure
and
the
degree
of
crystallinity
was
calculated
using
the
same
values
of
polynomial
order,
number
of
iteration
and
point
of
window.
Thermal
stability
of
native
A.
hippocastanum
starch
was
analyzed
by
TGA.
Fig.
3
shows
the
TGA
and
derivative
of
the
TG
curves
(DTG)
of
starch
in
the
temperature
range
of
20–900 C.
The
TGA/DTG
curves
revealed
three-steps
of
decomposition.
The
small
weight
loss
(around
7%)
between
30
and
140 C
was
attributed
to
the
evap-
oration
of
water
from
starch
granules.
The
percentage
of
weight
loss
in
this
step
was
similar
to
the
moisture
content
of
A.
hippocastanum
starch
determined
gravimetrically.
After
240 C
the
weight
loss
rate
quickly
increased
and
two
degradation
steps
were
distinguished
at
249
and
317 C
with
a
weight
loss
of
5.5%
and
38.4%,
respectively.
The
third
degradation
step
between
289
and
400 C
was
attributed
to
intermolecular
dehydration
to
levoglucosan
and
some
volatile
products,
such
as
carbon
dioxide,
lower
aldehydes,
methylfuranes,
and
ketones.
The
residue
at
900 C
of
around
10%,
probably
belongs
to
char
of
oxidized
organic
matter
and
inorganic
substances,
such
as
calcium,
sulfur
and
potassium,
detected
by
X-ray
fluorescence
spectrometry.
FTIR
spectrum
of
A.
hippocastanum
starch
in
the
2000–800
cm1
region
is
shown
in
Fig.
5.
Starches
from
different
botanical
sources
have
similar
FTIR
spectra
and
the
absorption
bands
have
been
assigned
in
many
previous
reports
(Cael
et
al.,
1975;
Santha,
Sudha,
Vijayakumari,
Nayar,
&
Moorthy,
1990).
The
broad
band
at
3416
cm1was
assigned
to
O–H
stretching
vibration
and
the
sharp
band
at
2930
cm1to
C–H
bond
stretching
of
methine
group
of
glucose
rings.
The
band
at
1646
cm1was
attributed
to
ı(O–H)
bending
vibration
of
water.
The
bands
at
1459
and
1419
cm1were
assigned
to
the
bending
vibrations
of
methylene
group
and
the
band
at
1369
cm1was
ascribed
to
C–H
symmetric
bending
vibration
of
methyl
group.
The
stretching
vibration
bands
of
C–C
and
C–O
appears
in
the
region
at
1300–800
cm1.
Their
intensity
has
been
related
to
amorphous
and
crystalline
phases
of
starch.
In
particu-
lar,
the
intensity
of
band
at
1019
cm1decreases
with
increasing
of
amorphous
state
(van
Soest,
Tournois,
de
Wit,
&
Vliegenthart,
1995).
3.2.
Characterization
of
TPS
from
A.
hippocastanum
starch
3.2.1.
Evaluation
of
processability
by
torque
rheometry
The
torque-rheometer
plots
as
a
function
of
time
showed
three
steps
for
TPS:
an
increasing
torque
from
the
beginning
until
a
maximum
value
is
reached,
attributed
to
a
plasticization
pro-
cess,
followed
by
a
stage
of
continuously
decreasing
values,
and
finally
the
steady
state.
The
torque
values
of
maximum
peak
and
steady
state,
and
the
plasticization
energy
are
summarized
in
Table
2.
TPS
prepared
from
A.
hippocastanum
starch
had
higher
plastifi-
cation
energies
and
maximum
torques
than
those
from
A.
araucana
starch.
The
type
of
plasticizer
employed
in
the
preparation
of
TPS
also
affected
the
values
of
plasticization
energy
and
torque
in
steady
state;
plasticization
energy
was
higher
for
glycerol:malic
acid
plasticized
TPS
than
for
glycerol
plasticized.
TPS
with
malic
acid
presented
the
higher
viscosity
(first
and
second
step)
due
to
the
reduction
of
mobility
of
the
polymeric
chains
as
a
result
of
chemical
modification
of
starch
by
malic
acid.
Olivato
et
al.
(2012a)
reported
that
multifunctional
organic
acids,
such
as
malic
acid,
were
able
to
interact
with
the
hydroxyl
groups
of
starch;
carboxyl
and
ester
groups
were
formed
during
reactive
extrusion,
promoting
esterifi-
cation
and
trans-esterification
reactions
of
starch.
Once
starch
granules
were
destructurized,
the
torque
decreased
until
reaching
a
steady
state.
CG-MA-30
and
PG-MA-30
samples
showed
the
lowest
torques
in
steady
state
(see
Table
2),
because
multifunctional
organic
acids
also
promote
the
acid
hydrolysis
of
the
glycosidic
linkages
of
starch,
decreasing
the
molecular
weight
of
amylose
and
amylopectin,
as
well
as
the
degree
of
polymerization
of
the
amylopectin
molecules
(Da
Róz,
Zambon,
Curvelo,
&
Carvalho,
2011).
Similar
behavior
of
rheological
properties
was
reported
by
Jiugao
et
al.
(2005)
for
glycerol
plasticized
starch
modified
with
citric
acid.
On
the
other
hand,
the
time
required
to
reach
steady
state
depended
on
the
type
of
plasticizer
and
type
of
starch
(see
Fig.
A2).
The
time
was
around
4
min
for
samples
without
malic
acid,
8
min,
for
the
CG-MA-30
sample,
and
11
min
for
the
PG-MA-30
sam-
ple.
Glycerol:malic
acid
plasticized
TPS
showed
a
slight
downward
trend
with
time
in
steady
state,
indicating
a
continuous
but
small
change
in
viscosity
of
these
samples.
The
reduction
of
viscosity
of
TPS
with
malic
acid
is
attributed
not
only
to
the
decrease
of
molec-
ular
weight
of
starch,
but
also
to
less
branch
chains
of
amylopectin
participating
in
the
formation
of
entanglements
structures
with
the
outer
chains
of
other
amylopectin,
or
with
amylose
(van
Soest
et
al.,
1996).
3.2.2.
Scanning
electron
microscopy
SEM
images
of
the
fractured
surfaces
(Fig.
6)
displayed
continu-
ous
phases.
However,
some
broken
starch
granules
were
detected
in
the
rough
fractured
surface
of
the
CG-30
sample
(Fig.
6A).
The
broken
granules
were
completely
embedded
in
the
TPS
matrix.
No
signs
of
granule
pullout
and
pits
on
the
surface,
and
no
obvious
682
J.
Casta˜
no
et
al.
/
Carbohydrate
Polymers
112
(2014)
677–685
Fig.
4.
TGA
and
DTG
curves
of
(A)
A.
hippocastanum
starch
and
(B)
TPS.
gaps
at
the
irregular
interfaces
between
the
broken
granule
and
the
matrix
were
observed.
Thus,
the
addition
of
30
wt.%
of
glycerol
in
the
CG-30
sample
was
not
enough
for
achieving
complete
plasti-
cization
of
TPS
under
used
processing
conditions.
Similar
results
were
reported
by
Qiao,
Tang,
and
Sun
(2011)
and
Mao,
Imam,
Gordon,
Cinelli,
and
Chiellini
(2000).
As
reported
previously
the
PG-30
sample
exhibited
a
rough
and
homogenous
surface
without
granules
(Casta˜
no
et
al.,
2012b).
CG-MA-30
and
PG-MA-30
samples
presented
smooth
and
homogenous
fractured
surfaces
with
some
randomly
distributed
holes.
Ma,
Chang,
Yu,
and
Stumborg
(2009)
reported
that
pits
found
in
fractured
surface
of
glycerol:citric
acid
plasticized
TPS
were
due
to
exfoliation
of
non-destructured
starch
granules.
However,
frag-
ments
or
entire
granules
were
not
observed
at
the
surface
of
these
samples,
as
expected
for
concentrations
of
organic
acids
higher
than
1
wt.%,
which
promote
the
disruption
and
dissolution
of
starch
granules
(Jiugao
et
al.,
2005).
3.2.3.
Thermal
stability
TGA
curves
of
TPS
exhibited
similar
behavior
with
three
steps
of
decomposition
(Fig.
4B).
Loss
weight
varied
slowly
from
room
tem-
perature
up
to
100 C,
with
a
weight
loss
of
about
2%.
The
maximum
decomposition
temperature
(Tmax)
and
associated
weight
loss
at
each
decomposition
step
depended
on
starch
source
and
plasticizer
used.
Glycerol:malic
acid
plasticized
TPS
exhibited
lower
thermal
stability
than
TPS
prepared
with
glycerol
alone,
due
to
a
reduction
of
molecular
weight
of
starch
chains
by
acid
hydrolysis
(Jiugao
et
al.,
2005).
Thermal
degradation
at
the
first
step
of
decomposition
was
asso-
ciated
to
loss
of
low
molecular
weight
substances
such
as
water
and
glycerol.
Tmax of
first
step
of
CG-30
sample
(176.9 C)
was
higher
than
that
of
PG-30
(156.9 C)
due
to
the
lower
water
content
of
A.
hippocastanum
(7.0
wt.%)
compared
to
A.
araucana
(8.1
wt.%).
With
decreasing
water
content
of
TPS
the
boiling
point
of
glycerol–water
mixture
increases.
This
behavior
is
more
pronounced
at
weight
fraction
of
water
lower
than
10
wt.%
in
water/glycerol
mix-
tures
(Chen
&
Thompson,
1970).
The
second
decomposition
peak
appeared
in
the
range
of
277–282 C
for
TPS
prepared
with
glycerol
and
in
the
range
of
257–262 C
for
glycerol:malic
acid
plasticized
TPS,
and
was
mainly
associated
to
evaporation
of
glycerol.
Note
that
weight
loss
at
Tmax of
second
peak
was
one-third
less
in
glycerol:malic
acid
plasticized
TPS.
After
300 C
the
weight
loss
cor-
responds
to
dehydration
and
decomposition
of
starch
(Liu,
Xie,
Yu,
Chen,
&
Li,
2009).
As
expected,
the
values
of
Tmax of
the
third
step
only
depended
on
starch
source,
being
301.9
and
306.9 C
for
TPS
prepared
with
A.
hippocastanum
starch
and
with
A.
araucana
starch,
respectively.
The
residual
weight
percentage
of
TPS
prepared
with
glyc-
erol:malic
acid
was
higher
than
those
prepared
with
glycerol.
Similar
results
were
reported
by
Shi
et
al.
(2007).
They
attributed
the
increasing
of
residual
weight
to
the
tendency
of
char
formation
of
esterified
starch.
3.2.4.
FTIR
analysis
The
changes
in
FTIR
spectra
between
2000
and
800
cm1of
A.
hippocastanum
starch
plasticized
with
glycerol
alone
and
with
glyc-
erol:malic
acid
by
melt
blending
are
shown
in
Fig.
4.
The
intensity
of
bands
between
1200
and
950
cm1decreased
in
TPS,
due
to
the
destructuration
and
decreasing
crystallinity
of
starch
(van
Soest
et
al.,
1995).
In
particular,
the
band
at
1019
cm1,
associated
to
C–O
bond
of
C–O–C
in
starch
(Jiugao
et
al.,
2005),
was
not
observed
in
CG-30
and
CG-MA-30
spectra.
However,
the
band
at
1052
cm1
(shoulder
in
FTIR
spectrum
of
A.
hippocastanum
starch)
was
shifted
to
a
lower
wave
numbers
in
CG-30
(1039
cm1)
and
CG-MA-30
(1035
cm1)
samples.
The
shift
of
band
at
1052
cm1to
lower
wave
numbers
(around
13
cm1)
in
TPS
prepared
with
glycerol
could
be
due
to
the
formation
of
hydrogen
bonds
among
C–O–H
and
C–O–C
groups
of
starch
and
OH
of
glycerol.
In
glycerol:malic
acid
plasti-
cized
TPS
this
shift
was
higher
(around
17
cm1),
probably
due
to
hydroxyl
and
carboxyl
groups
of
poly(carboxylic
acid)
interacting
stronger
with
C–O–H
and
C–O–C
groups
in
starch
than
with
glyc-
erol.
A
new
strong
band
appeared
at
1739
cm1on
FTIR
spectrum
of
CG-MA-30
sample,
which
corresponds
to
C
O
ester
carbonyl
group
stretching.
The
peak
at
1739
cm1is
associated
to
the
ester
bond
in
esterified
starch
(Diop,
Li,
Xie,
&
Shi,
2011;
Ma
et
al.,
2009).
3.2.5.
Tensile
properties
The
effect
of
starch
source
(A.
hippocastanum
and
A.
araucana)
on
the
tensile
properties
of
glycerol
plasticized
TPS
is
shown
in
Table
2,
where
tensile
modulus,
tensile
strength
and
elongation
at
break
of
TPS
are
summarized.
Note
that
tensile
modulus,
i.e.
tan-
gent
modulus,
was
calculated
at
10%,
30%
y
90%
of
strain,
because
the
samples
did
not
present
Hookean
behavior.
PG-30
compared
with
CG-30
exhibited
higher
values
of
tensile
properties.
Tensile
strength
was
60%
and
tensile
modulus
at
high
deformation
of
strain
was
almost
200%
higher;
the
rigidity
and
higher
strength
of
PG-30
was
associated
to
the
absence
of
residual
starch
granules
at
frac-
tured
surfaces
(Fig.
6B)
and
the
higher
apparent
amylose
content
of
A.
araucana
starch.
Chaudhary,
Torley,
Halley,
McCaffery,
and
Chaudhary
(2009)
studied
the
effect
of
amylose
content
on
tensile
properties
of
TPS
from
maize
obtained
by
melt
blending,
and
con-
cluded
that
high
amylose
TPS
tend
to
be
tougher
due
to
the
higher
J.
Casta˜
no
et
al.
/
Carbohydrate
Polymers
112
(2014)
677–685
683
Fig.
5.
FTIR
spectrum
of
A.
hippocastanum
starch
and
TPS.
Fig.
6.
SEM
micrographs
of
fracture
surfaces
of
molded
TPS
specimens:
(A)
CG-30;
(B)
PG-30;
(C)
CG-MA-30
and
(D)
PG-MA-30.
degree
of
interactions
among
the
double
helix
structure
of
amylose
and
the
outer
chains
of
amylopectin.
The
entanglement
between
amylose
and
amylopectin
gave
a
TPS
material
with
a
denser
poly-
mer
network.
Glycerol:malic
acid
plasticized
TPS
prepared
exhibited
a
sticky
behavior,
making
it
impossible
to
remove
the
specimens
from
the
mold
without
plastic
deformation.
The
sticky
behavior
was
attributed
to
the
fact
that
malic
acid
(15
wt.%)
promoted
acid
hydrolysis
of
the
glycosidic
linkages
in
starch.
Da
Róz
et
al.
(2011)
correlated
the
reduction
of
viscosity
and
the
stickiness
of
TPS
pre-
pared
with
glycerol:water:poly(carboxylic
acid)
with
the
decrease
of
molecular
weight
in
starch.
4.
Conclusion
A.
hippocastanum
L.
seed
is
an
attractive
source
of
non-edible
starch
for
the
production
of
starch-based
materials.
This
seed
is
easily
collected
and
presently
ends
up
in
landfills.
Starch
isolated
from
A.
hippocastanum
seed
shows
a
bimodal
size
distribution
with
characteristic
shapes
for
each
mode.
A.
hippocastanum
starch
dis-
played
a
typical
C-type
diffraction
pattern.
The
high
swelling
power
of
this
starch
is
due
to
the
proportion
of
short
branch
chains
of
amylopectin
(APC
ratio
=
0.396)
and
phosphorous
content
(0.012%).
Glycerol
plasticized
TPS
from
A.
hippocastanum
and
from
A.
aura-
cana
starch
showed
similar
rheological
and
thermal
properties.
684
J.
Casta˜
no
et
al.
/
Carbohydrate
Polymers
112
(2014)
677–685
Tensile
strength
and
young
modulus
were
lower
for
TPS
from
A.
hip-
pocastanum.
The
fractured
surface
of
glycerol
plasticized
TPS
from
A.
hippocastanum
starch
displayed
residual
starch
granules.
The
use
of
malic
acid
at
15
wt.%
as
co-plasticizer
adversely
affected
thermal
and
mechanical
properties
of
TPS
plasticized
by
glycerol.
Glycerol:malic
acid
plasticized
TPS
exhibited
pasty
and
sticky
behavior,
which
can
be
attributed
to
increasing
depolymer-
ization
of
starch
molecules
by
acid
hydrolysis
promoted
by
malic
acid.
Acknowledgments
This
work
has
been
financed
by
CONICYT-REGIONAL
R08C1002
and
Programa
de
Financiamiento
Basal
para
Centros
Científicos
y
Tecnológicos
de
Excelencia
PFB-27,
Projects
InnovaChile
13IDL2-
23120
and
Fondef
D09I1195.
Appendix
A.
See
Figs.
A1
and
A2.
Fig.
A1.
Storage
parenchyma
cells
from
cotyledon
of
mature
A.
hippocastanum
seed.
Fig.
A2.
Torque
variation
as
a
function
of
time
for
TPS.
References
Bertuzzi,
M.
A.,
Castro
Vidaurre,
E.
F.,
Armada,
M.,
&
Gottifredi,
J.
C.
(2007).
Water
vapor
permeability
of
edible
starch
based
films.
Journal
of
Food
Engineering,
80(3),
972–978.
Bharadwaj,
D.
N.
(2010).
Use
and
environment
impact
of
biodegradable
plastics
a
review.
Current
Advances
in
Agricultural
Sciences,
2(2),
65–69.
Cael,
J.
J.,
Koenig,
J.
L.,
&
Blackwell,
J.
(1975).
Infrared
and
Raman
spectroscopy
of
carbohydrates.
Part
VI:
Normal
coordinate
analysis
of
V-amylose.
Biopolymers,
14(9),
1885–1903.
Casta˜
no,
J.,
Bouza,
R.,
Rodríguez-Llamazares,
S.,
Carrasco,
C.,
&
Vinicius,
R.
V.
B.
(2012).
Processing
and
characterization
of
starch-based
materials
from
pehuen
seeds
(Araucaria
araucana
(Mol)
K.
Koch).
Carbohydrate
Polymers,
88(1),
299–307.
Casta˜
no,
J.,
Rodríguez-Llamazares,
S.,
Carrasco,
C.,
&
Bouza,
R.
(2012).
Physical,
chem-
ical
and
mechanical
properties
of
pehuen
cellulosic
husk
and
its
pehuen-starch
based
composites.
Carbohydrate
Polymers,
90,
1550–1556.
ˇ
Cukanovi´
c,
J.,
Nini´
c-Todorovi´
c,
J.,
Ognjanov,
V.,
Mladenovi´
c,
E.,
Ljubojevi´
c,
M.,
&
Kur-
jakov,
A.
(2011).
Biochemical
composition
of
the
horse
chestnut
seed
(Aesculus
hippocastanum
L.).
Archives
of
Biological
Sciences
Belgrade,
63(2),
345–351.
Chaudhary,
A.
L.,
Torley,
P.
J.,
Halley,
P.
J.,
McCaffery,
N.,
&
Chaudhary,
D.
S.
(2009).
Amylose
content
and
chemical
modification
effects
on
thermoplastic
starch
from
maize
Processing
and
characterisation
using
conventional
polymer
equipment.
Carbohydrate
Polymers,
78(4),
917–925.
Chen,
D.
H.
T.,
&
Thompson,
A.
R.
(1970).
Isobaric
vapor–liquid
equilibriums
for
the
systems
glycerol–water
and
glycerol–water
saturated
with
sodium
chloride.
Journal
of
Chemical
&
Engineering
Data,
15(4),
471–474.
Chung,
H.-J.,
&
Liu,
Q.
(2012).
Physicochemical
properties
and
in
vitro
digestibility
of
flour
and
starch
from
pea
(Pisum
sativum
L.)
cultivars.
International
Journal
of
Biological
Macromolecules,
50(1),
131–137.
Da
Róz,
A.
L.,
Zambon,
M.
D.,
Curvelo,
A.
A.
S.,
&
Carvalho,
A.
J.
F.
(2011).
Thermoplastic
starch
modified
during
melt
processing
with
organic
acids:
The
effect
of
molar
mass
on
thermal
and
mechanical
properties.
Industrial
Crops
and
Products,
33(1),
152–157.
Diop,
C.
I.
K.,
Li,
H.
L.,
Xie,
B.
J.,
&
Shi,
J.
(2011).
Effects
of
acetic
acid/acetic
anhy-
dride
ratios
on
the
properties
of
corn
starch
acetates.
Food
Chemistry,
126(4),
1662–1669.
Frost,
K.,
Kaminski,
D.,
Kirwan,
G.,
Lascaris,
E.,
&
Shanks,
R.
(2009).
Crystallinity
and
structure
of
starch
using
wide
angle
X-ray
scattering.
Carbohydrate
Polymers,
78(3),
543–548.
Han,
J.
H.
(2014).
Edible
films
and
coatings:
A
review.
In
J.
H.
Han
(Ed.),
Innovations
in
food
packaging
(2nd
ed.,
pp.
213–255).
San
Diego:
Academic
Press.
Hricovíniová,
Z.,
&
Babor,
K.
(1992).
Saccharide
constituents
of
horse
chestnut
(Aes-
culus
hippocastanum
L.)
seeds
II/isolation
and
characterization
of
the
starch.
Chemical
Papers,
45(3),
196–198.
Jane,
J.,
Chen,
Y.
Y.,
Lee,
L.
F.,
McPherson,
A.
E.,
Wong,
K.
S.,
Radosavljevic,
M.,
et
al.
(1999).
Effects
of
amylopectin
branch
chain
length
and
amylose
content
on
the
gelatinization
and
pasting
properties
of
starch.
Cereal
Chemistry
Journal,
76(5),
629–637.
Jane,
J.-L.
(2009).
Chemistry
and
technology.
In
J.
BeMiller,
&
R.
Whistler
(Eds.),
Structural
features
of
starch
granules
II
(3rd
ed.,
pp.
205–225).
Jiugao,
Y.,
Ning,
W.,
&
Xiaofei,
M.
(2005).
The
effects
of
citric
acid
on
the
properties
of
thermoplastic
starch
plasticized
by
glycerol.
Starch
Stärke,
57(10),
494–504.
Liu,
H.,
Xie,
F.,
Yu,
L.,
Chen,
L.,
&
Li,
L.
(2009).
Thermal
processing
of
starch-based
polymers.
Progress
in
Polymer
Science,
34(12),
1348–1368.
Ma,
X.,
Chang,
P.
R.,
Yu,
J.,
&
Stumborg,
M.
(2009).
Properties
of
biodegradable
citric
acid-modified
granular
starch/thermoplastic
pea
starch
composites.
Carbohy-
drate
Polymers,
75(1),
1–8.
Mao,
L.,
Imam,
S.,
Gordon,
S.,
Cinelli,
P.,
&
Chiellini,
E.
(2000).
Extruded
cornstarch–glycerol–polyvinyl
alcohol
blends:
Mechanical
properties,
mor-
phology,
and
biodegradability.
Journal
of
Polymers
and
the
Environment,
8(4),
205–211.
Moraes,
J.,
Alves,
F.
S.,
&
Franco,
C.
M.
L.
(2013).
Effect
of
ball
milling
on
structural
and
physicochemical
characteristics
of
cassava
and
Peruvian
carrot
starches.
Starch
Stärke,
65(3–4),
200–209.
Morán,
J.
I.,
Vázquez,
A.,
&
Cyras,
V.
P.
(2013).
Bio-nanocomposites
based
on
deriva-
tized
potato
starch
and
cellulose,
preparation
and
characterization.
Journal
of
Materials
Science,
48(20),
7196–7203.
Musatenko,
L.
I.,
Generalova,
V.
N.,
Martyn,
G.
I.,
Vedenicheva,
N.
P.,
&
Vasyuk,
V.
A.
(2003).
Hormonal
complex
and
ultrastructure
of
maturing
Aesculus
hippocas-
tanum
seeds.
Russian
Journal
of
Plant
Physiology,
50(3),
360–364.
Olivato,
J.
B.,
Grossmann,
M.
V.
E.,
Bilck,
A.
P.,
&
Yamashita,
F.
(2012).
Effect
of
organic
acids
as
additives
on
the
performance
of
thermoplastic
starch/polyester
blown
films.
Carbohydrate
Polymers,
90(1),
159–164.
Olivato,
J.
B.,
Grossmann,
M.
V.
E.,
Yamashita,
F.,
Eiras,
D.,
&
Pessan,
L.
A.
(2012).
Citric
acid
and
maleic
anhydride
as
compatibilizers
in
starch/poly(butylene
adipate-
co-terephthalate)
blends
by
one-step
reactive
extrusion.
Carbohydrate
Polymers,
87(4),
2614–2618.
Qiao,
X.,
Tang,
Z.,
&
Sun,
K.
(2011).
Plasticization
of
corn
starch
by
polyol
mixtures.
Carbohydrate
Polymers,
83,
659–664.
Ruxanda,
B.,
&
Teac ´
ˆ
a,
C.
A.
(2012).
Effects
of
chemical
modification
on
the
structure
and
mechanical
properties
of
starch-based
biofilms.
Monatshefte
für
Chemie,
143,
335–343.
Santha,
N.,
Sudha,
K.
G.,
Vijayakumari,
K.
P.,
Nayar,
V.
U.,
&
Moorthy,
S.
N.
(1990).
Raman
and
infrared
spectra
of
starch
samples
of
sweet
potato
and
cassava.
Journal
of
Chemical
Sciences,
102(5),
705–712.
J.
Casta˜
no
et
al.
/
Carbohydrate
Polymers
112
(2014)
677–685
685
Shi,
R.,
Zhang,
Z.,
Liu,
Q.,
Han,
Y.,
Zhang,
L.,
Chen,
D.,
et
al.
(2007).
Characteriza-
tion
of
citric
acid/glycerol
co-plasticized
thermoplastic
starch
prepared
by
melt
blending.
Carbohydrate
Polymers,
69(4),
748–755.
Singh,
J.,
McCarthy,
O.
J.,
Singh,
H.,
Moughan,
P.
J.,
&
Kaur,
L.
(2007).
Morphological,
thermal
and
rheological
characterization
of
starch
isolated
from
New
Zealand
Kamo
Kamo
(Cucurbita
pepo)
fruit
A
novel
source.
Carbohydrate
Polymers,
67(2),
233–244.
Singh,
N.,
Kaur,
S.,
Isono,
N.,
Ichihashi,
Y.,
Noda,
T.,
Kaur,
A.,
et
al.
(2012).
Diver-
sity
in
characteristics
of
starch
amongst
rice
bean
(Vigna
umbellate)
germplasm:
Amylopectin
structure,
granules
size
distribution,
thermal
and
rheology.
Food
Research
International,
46(1),
194–200.
Soni,
P.
L.,
Sharma,
H.
W.,
&
Bisen,
S.
S.
(1989).
Physicochemical
properties
of
Careya
arborea
and
Aesculus
assamica
starches.
Starch
Stärke,
41(6),
208–211.
Srichuwong,
S.,
Sunarti,
T.
C.,
Mishima,
T.,
Isono,
N.,
&
Hisamatsu,
M.
(2005a).
Starches
from
different
botanical
sources
I:
Contribution
of
amylopectin
fine
structure
to
thermal
properties
and
enzyme
digestibility.
Carbohydrate
Polymers,
60(4),
529–538.
Srichuwong,
S.,
Sunarti,
T.
C.,
Mishima,
T.,
Isono,
N.,
&
Hisamatsu,
M.
(2005b).
Starches
from
different
botanical
sources
II:
Contribution
of
starch
structure
to
swelling
and
pasting
properties.
Carbohydrate
Polymers,
62(1),
25–34.
van
Soest,
J.
J.
G.,
Benes,
K.,
de
Wit,
D.,
&
Vliegenthart,
J.
F.
G.
(1996).
The
influence
of
starch
molecular
mass
on
the
properties
of
extruded
thermoplastic
starch.
Polymer,
37(16),
3543–3552.
van
Soest,
J.
J.
G.,
Tournois,
H.,
de
Wit,
D.,
&
Vliegenthart,
J.
F.
G.
(1995).
Short-range
structure
in
(partially)
crystalline
potato
starch
determined
with
attenuated
total
reflectance
Fourier-transform
IR
spectroscopy.
Carbohydrate
Research,
279,
201–214.
Vieira,
M.,
da
Silva,
M.,
dos
Santos,
L.,
&
Beppu,
M.
(2011).
Natural-based
plas-
ticizers
and
biopolymer
films:
A
review.
European
Polymer
Journal,
47(3),
254–263.
Wani,
I.
A.,
Jabeen,
M.,
Geelani,
H.,
Masoodi,
F.
A.,
Saba,
I.,
&
Muzaffar,
S.
(2014).
Effect
of
gamma
irradiation
on
physicochemical
properties
of
Indian
Horse
Chestnut
(Aesculus
indica
Colebr.)
starch.
Food
Hydrocolloids,
35,
253–263.
Wong,
K.
S.,
&
Jane,
J.
(1995).
Effects
of
pushing
agents
on
the
separation
and
detection
of
debranched
amylopectin
by
high-performance
anion-exchange
chromatography
with
pulsed
amperometric
detection.
Journal
of
Liquid
Chro-
matography,
18(1),
63–80.
Zullo,
R.,
&
Iannace,
S.
(2009).
The
effects
of
different
starch
sources
and
plasticizers
on
film
blowing
of
thermoplastic
starch:
Correlation
among
process,
elonga-
tional
properties
and
macromolecular
structure.
Carbohydrate
Polymers,
77(2),
376–383.
Zuraida,
A.,
Yusliza,
Y.,
Anuar,
H.,
&
Mohd
Khairul
Muhaimin,
R.
(2012).
The
effect
of
water
and
citric
acid
on
sago
starch
bio-plastics.
International
Food
Research
Journal,
19(2),
715–719.
... Thermoplastic starch (TPS) is a relatively new substance for use as biodegradable plastic, and it is now one of the most researched polymers in this sector [1]. TPS is made by breaking the crystal structure of starch with a plasticizer at high temperatures, both with and without shear stress [2]. There have been multiple examples of starch made for TPS from edible fruits, vegetables, seeds, and other agricultural products [3]. ...
... There have been numerous non-edible sources of starch discovered, including the horse chestnut (Aesculus hippocastanum L.), which contains more than 35% starch by weight [5]. Many countries, including England, Chile, India and others, have horse chestnut available locally [2,6]. Horse chestnut seeds are commonly used for decoration, yet the vast majority of them end up in landfills. ...
... The preparation of thermoplastic from horse chestnut starch was reported by CAS-TAñO et al. [2]. The authors state that the plasticizers were combined with the starch prior to melt-blending, and that the mixture was then blended for 15 min at 120°C and 60 rpm. ...
Chapter
The starch-based materials such as thermoplastic starch film are a promising alternative to non-renewable petroleum-based plastics. The development of an alternative conventional plastic from bio-based materials has gained great interest following its biodegradable, non-hazardous and renewable advantages. Following that, horse chestnut is an exciting source of starch for producing thermoplastic starch film. Nonetheless, the thermoplastic starch film is weak in strength and easily affected by water due to its highly hydrophilic property, thus limiting its practicability. In this regard, the additions of nanocellulose into thermoplastic starch have shown drastic improvement in its mechanical properties and water permeability of the film. This chapter discusses the potential of nanocellulose reinforced plasticized starch from horse chestnut as a replacement for petroleum-based plastic in packaging applications.
... Pea starch, PeS, with an amylose content of 35% according to the manufacturer, was kindly provided by Roquette Freres S.A. (Lestrem, France). Starches of ripe Aesculus hippocastanum, AeS, 16 and Araucaria araucana, ArS, 17 seeds, with an amylose content of 33% and 28%, respectively, isolated by the wet milling method, were previously characterized by us. Gallic acid (GA), ABTS • (2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid)), Trolox (6-hydroxy-2,5,7,8-tetramethyl chroman-2-carboxylic acid), and TPTZ (2,4,6-tris(2-pyridyl)-s-triazine (TPTZ) were purchased from Sigma-Aldrich S.A. (Santiago, Chile). ...
... Swelling Power and Solubility. The swelling power and solubility of native starches and unloaded and loaded starch microparticles were determined at 60°C using the method reported by Castanõ et al. 16 with modification. Briefly, an aqueous suspension of 3.3% w/v native starch or starch microparticle was prepared using water preheating at 30°C and then was heated to 60°C in water bath for 30 min. ...
... 14 AeS exhibited, in addition to the typical peaks of C-type allomorph, two shoulder peaks at about 22°a nd 24°(2θ). 16 3.2. Characterization of Loaded and Unloaded Starch Microparticles. ...
Article
Donut-shaped microparticles were prepared from starches of different botanical sources by heat-mixture-alcoholic treatment. C-type starch allomorphs: pea, Araucaria araucana, and Aesculus hippocastanum were used for the preparation of the microparticles. The microparticles were characterized and evaluated as a carrier of phenolic compounds. The gallic acid encapsulation efficiency was determined, and its antioxidant activity was evaluated using spectrophotometric assays. The starch microparticles exhibited a round to elongated shape, with a central concavity, similar to donut-shaped microparticles. In the loaded and unloaded starch microparticles, B-type and V-type polymorphs coexist. The entrapment of gallic acid in starch microparticles is due to hydrogen-bond formation and not to an amylose inclusion complex. The loaded pea starch microparticles showed lower onset and peak gelatinization temperature with respect to their unloaded counterpart, which is associated with the decreased stability of formed crystallites. The percentage of phenolic compounds encapsulated depended on the native starch precursor. Pea starch microparticles loaded with gallic acid showed the highest encapsulation efficiency around 45%. Gallic acid entrapped in starch microparticles showed no loss of antioxidant activity. Starch microparticles are a food-grade solid matrix suitable as carriers for active compounds.
... Carrageenan is frequently used in the food industry in a variety of products due to its gelling, stabilizing, and thickening properties, as well as its capacity to form thermo-reversible gels (Castaño et al., 2017). The combination of starch and carrageenan is utilized to create biofilms that have a number of benefits over those created using synthetic material (Castaño et al., 2014;Galus & Kadzińska, 2016). ...
Article
Full-text available
Microbial fuel cells (MFCs) are an innovative method that generates sustainable electricity by exploiting the metabolic processes of microorganisms. The membrane that divides the anode and cathode chambers is an important component of MFCs. Commercially available membranes, such as Nafion, are both costly, not sustainable, and harmful to the environment. In this study, a low-cost alternative membrane for MFCs based on a starch-carrageenan blend (SCB-LCM) was synthesized. The SCB-LCM membrane was created by combining starch and carrageenan and demonstrated a high dehydration rate of 98.87 % over six hours. SEM analysis revealed a smooth surface morphology with no pores on the membrane surface. The performance of SCB-LCM membrane-based MFCs was evaluated and compared to that of other membranes, including Nafion 117 and Nafion 212. All membranes tested over 25 hours lost significant weight, with SCB-LCM losing the least. The maximum power density (MPD) of the SCB-LCM MFCs was 15.77 ± 4.34 mW/m2, indicating comparable performance to commercial membranes. Moreover, the cost-to-power ratio for MFCs employing SCB-LCM was the lowest (0.03 USD.m2/mW) when compared to other membranes, indicating that SCB-LCM might be a viable and cost-effective alternative to Nafion in MFCs. These SCB-LCM findings lay the groundwork for future research into low-cost and sustainable membrane for MFC technologies.
... In addition, the SEM images demonstrated that the fibers after alkaline treatment have micrometric dimensions. Figure 6b (left) shows a SEM image for corn starch, where the presence of grains characteristic of corn starch is observed [62][63][64]. However, the images for the TPS films displayed in Figure 6b,c reveal the loss of a large part of the starch grains; this behavior occurred due to plastification with glycerol, which destroys the starch grains [18,51]. ...
Article
Full-text available
Citation: de Vilhena, M.B.; Matos, R.M.; Ramos Junior, G.S.d.S.; Viegas, B.M.; da Silva Junior, C.A.B.; Macedo, E.N.; Paula, M.V.d.S.; da Silva Souza, J.A.; Candido, V.S.; de Sousa Cunha, E.J. Influence of Glycerol and SISAL Microfiber Contents on the Thermal and Tensile Properties of Thermoplastic Starch Composites. Polymers 2023, 15, 4141. https:// Abstract: The increasing use of petroleum plastics has caused environmental damage due to the degradation time of these materials. An alternative to petroleum plastics could be thermoplastic starch (TPS). However, thermoplastic starch does not exhibit satisfactory tensile properties. The mechanical properties of thermoplastic starch can be improved by adding sisal microfibers. Thus, the objective of this study was to evaluate the influence of different levels of glycerol and sisal microfibers on the thermal and tensile properties of thermoplastic corn starch composites. The microfibers were obtained via mechanical treatment followed by chemical treatment (alkaline treatment and bleaching). The films were obtained by the casting method using commercial corn starch and glycerol as a plasticizing agent, reinforced with sisal microfibers. Fourier transform infrared spectroscopy (FTIR) results revealed that the addition of microfibers did not change the chemical structure of the TPS matrix. The films from the samples with 18% glycerol and 10% microfibers had the highest value for the maximum tension, equal to 4.78 MPa. The thermal decomposition profile of TPS was not altered by the addition of microfibers. Our findings demonstrated the profound influence of glycerol and microfiber contents on the tensile properties of thermoplastic starch composites.
Article
The most widely cultivated varieties of potato in Chile are Patagonia and Asterix. However, there is scarce information about the physical–chemical characteristics of starches isolated from these potatoes. In this study, the starches from two potato varieties, Patagonia and Asterix, were isolated and characterised. In addition, an exploratory study of the preparation of starch nanoparticles from these potato varieties by the flash nanoprecipitation method was performed. The morphology (scanning electron microscopy and optical microscopy), crystallinity (X-ray diffraction), thermal (differential scanning calorimetry), rheological (rapid visco analyser) and hydration properties are reported. Starches from Patagonia potato (PPS) and Asterix potato (APS) were isolated by the wet milling method. PPS and APS had similar starch content on fresh weight basis, around 15%. The granule particle sizes of Chilean potato starches showed a unimodal distribution, in the range of 10–90 µm. Amylose content varied between 23.8 and 25.3% for APS and PPS, respectively, making them normal starches. Gelatinisation temperature and enthalpy were 58.6 and 60.1 °C and 15.5 and 17.8 J/g for APS and PPS, respectively. Significant differences were obtained in the pasting parameters of Chilean potato starches; the rapid visco analyser pasting profiles of PPS exhibited a large viscosity peak and APS did not show a distinctive viscosity peak. The pasting profiles of APS are required for food products such as soups and sauces. Spherical nanoparticles were prepared from the Chilean potato starches by flash nanoprecipitation method using two micromixer devices: confined impinging jet mixer and multi-inlet vortex mixer. The size of potato nanoparticles depended on micromixer device, and the lowest particle size (100 nm) was obtained when the confined impinging jet mixer was used. The nanoparticles from Chilean potato starches are potential raw materials for the nutraceutical, food and pharmaceutical fields.
Article
Growing public concern toward environmental sustainability is currently motivating a paradigm shift toward designing easily degradable plastics that can replace conventional synthetic plastics. Teh massive rise in food waste generation TEMPhas led to an increased burden on landfills, thereby resulting in teh higher emission of greenhouse gases. Using dis food waste to produce bioplastics will benefit not only teh environment but also develop a systematic food waste management system. Moreover, bioplastics are preferred due to teh use of biomaterials derived from renewable resources. Furthermore, bioplastics degrade faster TEMPthan conventional synthetic plastics, which take years to degrade. Teh biodegradation of bioplastics occurs under normal environmental conditions and disintegrates into carbon dioxide, water, biomass, and inorganic compounds without producing hazardous residues. In dis review, we will discuss teh synthesis of starch based bioplastics using discarded parts of various fruits and vegetables. Furthermore, we will address teh importance of various components in teh development of starch based bioplastics, such as fillers, plasticizers, and other additives that are essential in providing teh bioplastic with different physio-mechanical properties. Therefore, bioplastic production using food waste will pave teh way to achieve systematic waste management and environmental sustainability in teh near future.
Article
A mild approach for synthesis of Cu2O nanoparticles by using pectin were described through a simple and ‘green’ way under ultrasonic conditions. The polar functional groups on the pectin (OH) facilitate the capping and stabilization of Cu2O nanoparticles. Structural features of the Cu2O NPs/pectin were assessed over several advanced techniques like UV-Vis spectroscopy, scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDS), energy dispersive X-ray analysis (EDX) elemental mapping, transmission electron microscopy (TEM), X-ray diffraction analysis (XRD), and inductively coupled plasma-atomic emission spectrometry (ICP-AES) analysis. It was catalytically explored in reducing an organic dye (Methylene blue-MB) in the presence of NaBH4 at mild conditions. The Cu2O NPs/pectin was recycled 9 times keeping consistency in its reactivity. IC50 values of the nanocomposite were found at 60, 54 and 53 µg/mL against CL40, SW1417, and CW2 cell lines respectively and accordingly, CW2 afforded the best carcinoma activity. The nanocomposite exhibited excellent cytotoxicity against colorectal carcinoma (CL40, SW1417, and CW2) cell lines without affecting the normal (HUVEC) cell line.
Article
A unique starch encapsulated Cu2O nanoparticles were synthesized through a simple and ‘green’ route using ultrasonic irradiation. The polar functional groups on the starch (OH) facilitate the NP capping and stabilization. Structural features of the material were assessed over several advanced techniques like fourier transformed infra-red (FT-IR), UV-Vis spectroscopy, scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDS), energy dispersive X-ray analysis (EDX) elemental mapping, transmission electron microscopy (TEM), X-ray diffraction analysis (XRD), and inductively coupled plasma-atomic emission spectrometry (ICP-AES) analysis. It was catalytically explored in reducing an organic dye (Methylene blue - MB) in the presence of NaBH4 at ambient conditions, being monitored in a UV-Vis spectrophotometer. The nanocatalyst was recycled 11 times keeping consistency in its reactivity. Biologically, the nanocomposite exhibited excellent cytotoxicity against lung adenocarcinoma (PC-14, LC-2/ad and HLC-1) cell lines without affecting the normal (HUVEC) cell line. IC50values of the nanocomposite were found at 618, 56 and 379 against HLC-1, LC-2/ad, and PC-14 cell lines respectively and accordingly, PC-14 afforded the best adenocarcinoma activity.
Article
Full-text available
Species of the genus Aesculus are very attractive woody ornamentals. Their organs contain numerous health‑promoting phytochemicals. The most valuable of them—aescin—is used in commercial preparations for the treatment of venous insufficiency. The industrial source of aescin is horse chestnut seeds because the zygotic embryos are the main site of its accumulation. Horse chestnut somatic and zygotic embryos contain similar amount of aescin, hence somatic embryos could be exploited as an alternative source of aescin. Somatic embryogenesis, androgenesis and de novo shoot organogenesis were successfully achieved in several Aesculus species, as well as secondary somatic embryogenesis and shoot organogenesis, which enables mass production of embryos and shoots. In addition, an efficient method for cryopreservation of embryogenic tissue was established, assuring constant availability of the plant material. The developed methods are suitable for clonal propagation of elite specimens selected as the best aescin producers, the most attractive ornamentals or plants resistant to pests and diseases. These methods are also useful for molecular breeding purposes. Thus, in this review, the medicinal uses and a comprehensive survey of in vitro propagation methods established for Aesculus species, as well as the feasibility of in vitro production of aescin, are presented and discussed.
Article
Full-text available
Starch is a biodegradable polymer produced in abundance from many renewable resources. This study examined the influence of citric acid (0-40% w/wt%)) and water (0-40% w/wt%) as secondary additive and glycerol as plasticizer on the mechanical properties of bio-plastic starch (BPS) from Malaysian sago. The CA content varies from 0 to 40 w/wt% while water was also varied from 0 to 40 w/wt%. FT-IR spectroscopy showed that acid citric improve the properties of BPS and water give negative effects to the carbon hydrogenbond. It is obvious that the addition of the CA at 30 wt/wt% improve the mechanical property of BPS to more than 40% compare to the addition of water.
Article
Full-text available
Starch isolated from the seeds of horse chestnut was distributed into fine-grain and coarse-grain portions. Amylopectin was separated and the ß-limit dextrin prepared was character-ized by periodate oxidation. The a-and ß-amylase susceptibilities were measured; it is de-duced that the horse chestnut starch possesses a lower amylose content, short outer and longer inner chains than those of the reference potato starch. Horse chestnut seeds represent — after extrac-tion of triterpenoids with alcohol — an untraditional source of saccharides: in addition to L-arabinose and r>glucuronic acid ethanol can be obtained from the fermentable saccharides [1]. This paper describes the isolation and characterization of starch constituting a dominant portion of the above-mentioned horse chestnut debris. The debris was filtered through a miller's cloth, floated and alternatively also extracted with di-methyl sulfoxide to afford the starch. Isolation of starch by extraction from the pasty starting ma-terial with perchloric acid followed by precipi-tation with iodine proved unsuccessful on the preparative scale. Washing of the ground mate-rial through a fine miller's cloth led to a high yield of starch (ca. 70 % of the starting material), which was partly contaminated by a fibrous material. Successive sedimentation of the starch was as-sociated with a concurrent separation of grains according to their size; thus, fine-grain and coarse-grain starches were obtained. On the other hand, extraction with dimethyl sulfoxide afforded a more pure starch, but in a substantially lower yield, so that this material cannot be considered a repre-sentative sample. Isolation of starch from the seeds was more recently described [2] as diffi-cult, because common procedures do not suffi-ciently attack the plasmatic substance needed for liberating starch from the cells; also the starch grains were then supposed as ones of the finest. Fractions of starch obtained from the debris of horse chestnut were checked for purity and yield by sorption of iodine. This method employed for * For Part / see Ref. [1].
Article
Full-text available
The use of native starch as a thermoplastic polymer is limited by its fragility and high moisture absorption. Native potato starch was derivatized using different agents in order to reduce its hydrophobicity. A reduction in starch mechanical properties was also found as a side effect of derivatization treatments. The addition of nanocellulose allows to considerably improve mechanical properties of biopolymers as well as to maintain their inherent biodegradation capability. Cellulose nanofibers were added to native and modified starch in order to improve the mechanical properties. The mechanical properties of starch/cellulose nanocomposite films were evaluated, finding that they considerably improved with the addition of cellulosic nanofibers. This contribution demonstrates a new approach in order to produce bioplastics with combined effect: improved moisture and heat resistance (due to the chemical treatment) and high mechanical properties (after the addition of nanocellulose).
Article
Full-text available
In this work, chemically modified corn starch and plasticized corn starch biofilms were obtained and characterized in four steps: (1) preparation of corn starch microparticles, (2) preparation of malic acid-modified corn starch microparticles (MA–SM), (3) preparation of corn starch biofilms and MA–SM-plasticized corn starch biofilms, and (4) characterization of the biofilms. The effects of MA–SM concentration (4, 8, and 12% based on the amount of corn starch) on the structural characteristics and mechanical properties of the biofilms were investigated. Changes in the starch granules after chemical modification were studied by X-ray diffraction, FT-IR spectroscopy, and scanning electron microscopy. The presence of ester carbonyl group stretching vibration at 1,720 cm−1 in FT-IR spectra was evidence of reaction of the starch microparticles with malic acid. The tensile yield strength and Young’s modulus of the films increased with increasing MA–SM content. Water uptake decreased from 69.8% for biofilm without MA–SM to 52.7% for biofilm with MA–SM. The improvement of these properties in the plasticized product could be attributed to the good interaction between the MA–SM filler and the corn starch. Graphical Abstract
Article
Edible films and coatings are produced from edible biopolymers and food-grade additives. Film-forming biopolymers can be proteins, polysaccharides (carbohydrates and gums), lipids, or a mixture of these Gennadios et al., 1997 (see Figure 9.1). Plasticizers and other additives are combined with the film-forming biopolymers to modify the physical properties or other functionality of the edible films. Biopolymers have multiple film-forming mechanisms, including intermolecular forces such as covalent bonds (e.g., disulfide bonds and crosslinking) and electrostatic, hydrophobic, or ionic interactions. For the resulting films or coatings to be edible, the film-forming mechanism involved in fabrication should be an appropriate food process-namely, pH modification, salt addition, heating, enzymatic modification, drying, use of food-grade solvents, or reactions with other food-grade chemicals. The control of fabrication process conditions is very important because the changes in treatment conditions can alter kinetics and reaction mechanisms (Guilbert et al., 1996, 1997).
Article
Starch isolated from Indian Horse Chestnut (Aesculus indica Colebr.) was subject to irradiation at 0, 5, 10, 15 kGy doses. Effect of irradiation on physicochemical properties of native starch was studied. The result revealed increase in water absorption capacity from 0.94 to 1.00 g/g, carboxyl content from 0.00 to 0.06%, solubility from 0.15 to 0.53 g/g and freeze thaw stability. Syneresis, pasting properties and pH were reduced following irradiation treatment. Syneresis decreased from 3.47 to 0.64% after 120 h refrigerated storage. Peak viscosity reduced from 5156.5 to 1422.5 cP, setback viscosity from 1191.5 to 73.0 cP and final viscosity from 3232.0 to 410.5 cP. X-ray diffraction pattern showed A type of pattern in native as well as irradiated starches. Granule morphology of native and irradiated starches under scanning electron microscope revealed that granules were round, oval, irregular or elliptical with smooth surfaces. Pearson correlation studies revealed that irradiation dose was positively correlated with water absorption capacity, oil absorption capacity, and solubility index and negatively correlated with syneresis, swelling index, freeze thaw and pasting properties.
Article
Peruvian carrot and cassava starches were ground in a ball mill for 4, 8, 16, and 32 h and their structural and physicochemical characteristics were determined. Results obtained from HPAEC-PAD, GPC, and amylose content indicated a breaking of hydrogen bounds and α-(1 → 6) linkages of the starch molecules after treatment. X-ray diffractograms showed that the milling provided a reduction in the crystalline area of the starch granules. Most of the starch granules displayed agglomeration after 4 h of milling, when observed under a scanning electron microscope, and after 16 h a shapeless mass was observed for Peruvian carrot starch. Solubility and water absorption capacity of the starches increased with an increase in the milling time, while RVA profiles showed a progressive reduction of peak, breakdown, and final viscosities, as well as the development of initial viscosity. Gelatinization temperatures and enthalpies were reduced. Prolonged ball milling accelerated the enthalpy relaxation in both starches. These results confirmed a partial gelatinization of the starches, which was 82.6% for Peruvian carrot and 65.4% for cassava starches after 32 h of milling. The Peruvian carrot starch was more affected by the ball milling because of both its lower amylose content and the defects in its crystalline structure.