You are on page 1of 24

Accepted Manuscript

Potato Fiber modified Thermoplastic Starch: Effects of Fiber Content on Ma-


terial Properties and Compound Characteristics

Barbara Fahrngruber, Johanna Eichelter, Sabine Erhäusl, Bernhard Seidl,


Rupert Wimmer, Norbert Mundigler

PII: S0014-3057(18)31676-8
DOI: https://doi.org/10.1016/j.eurpolymj.2018.10.050
Reference: EPJ 8683

To appear in: European Polymer Journal

Received Date: 3 September 2018


Revised Date: 29 October 2018
Accepted Date: 31 October 2018

Please cite this article as: Fahrngruber, B., Eichelter, J., Erhäusl, S., Seidl, B., Wimmer, R., Mundigler, N., Potato
Fiber modified Thermoplastic Starch: Effects of Fiber Content on Material Properties and Compound
Characteristics, European Polymer Journal (2018), doi: https://doi.org/10.1016/j.eurpolymj.2018.10.050

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Potato Fiber modified Thermoplastic Starch: Effects
of Fiber Content on Material Properties and
Compound Characteristics
a, b a a c d
Barbara Fahrngruber *, Johanna Eichelter , Sabine Erhäusl , Bernhard Seidl , Rupert Wimmer , Norbert Mundigler
a
Agrana Research & Innovation Center GmbH, Josef-Reither-Straße 21-23, 3430 Tulln, Austria
b
Polymer and Composite Engineering (PaCE) Group, Institute of Materials Chemistry and Research, Faculty of Chemistry, Universi ty of
Vienna, Währingerstraße 42, 1090 Vienna, Austria - Laboratory of Polymer Engineering LKT-TGM, Wexstraße 19-23, 1200 Vienna, Austria
c
Institute for Wood Technology and Renewable Materials, University for Natural Resources and Life Sciences, Vienna, Konrad Lor enz
Strasse 24, 3430 Tulln an der Donau, Austria
d
Institute for Natural Materials Technology, University for Natural Resources and Life Sciences, Vienna, IFA Tulln, Konrad Lor enz Strasse 20,
3430 Tulln an der Donau, Austria
* Corresponding author. Tel.: +43 2272 602 11410; Fax: +43 2272 602 11420. E-mail: barbara.fahrngruber@agrana.com (B. Fahrngruber).

The implementation of bio-based additives in polymer formulations provides possibilities to tailor material
characteristics for more sustainable solutions. Fibers derived from natural sources and of sustainable origin have
been gaining importance for bio-based as well as biodegradable packaging, showing also high performance. In
this research, potato fiber modified thermoplastic starch (1, 3 and 5 wt.% fiber content) was prepared, which was
used in the subsequent production of TPS/polyester compounds. Characterization results revealed a reduced
formation of closely-packed crystalline structures in TPS, along with increasing fiber content. Likewise, an
increase in materials rigidity was observed, as determined via dynamic mechanical analysis. Within the
compounded material, potato fiber additions enhanced a molecular weight reduction in starch while the
formation of agglomerates was induced. The prepared films exhibited increased mechanical stiffness and also
modified sorption characteristics. The identified property profiles are instrumental in creating novel, custom-
tailored packaging materials.

1. Introduction

It is a common approach to prepare compounds consisting of thermoplastic starch (TPS) and


biopolyesters (e.g. poly(butylene adipate-co-terephthalate) (PBAT)), as these materials are
inherently biodegradable, allowing also a shaping and forming with conventional polymer
processes such as extrusion and injection molding. As for the TPS-compounds, starch is
known to be limiting the mechanical material properties [1-5]. To a certain extent, properties
such as tensile strength and elongation at break can be adjusted through plasticizers and
additives of different types and concentrations [6-8]. However, the effectiveness of
plasticizers and additives has limitations, as they could leach out during subsequent
processing, or when in contact with e.g. food during storage. Increasing the mechanical
characteristics of TPS-based packaging films is a desired goal that might be better achievable

1
by e.g. an incorporation of fibrous fillers [9-11]. Due to their apparent hydrophilic and
hydrophobic functionalities, mixtures of TPS and polyester exhibit poor phase compatibility
[12]. As a consequence, the influence of fibrous additions on the miscibility with the matrix
polymer, in connection with achievable interfacial modifications, need to be critically
assessed. Therefore, the overall goal of this research is the assessment of co-stream derived
potato fibers utilized as a novel filler component in TPS. In the food industry, potato fibers
are known for their non-digestible dietary purposes, and they consist chemically of insoluble
fiber components such as cellulose, hemicelluloses and lignin, and of soluble components
such as pectin [13, 14]. Here, potato fibers are added to TPS, with the latter used in the
TPS/PBAT compound production. The effects of potato fibers on structural characteristics at
the macro and micro scale, along with mechanical and sorption properties of the obtained
materials, are assessed. This could be critical for functional packaging applications, where
specific material characteristics are required to ensure sufficient protection of packed goods.
The following hypotheses are stated:
- As additives and fillers usually change mixing conditions during extrusion, it is
hypothesized that potato fiber additions enhance shear forces, which potentially alter
molecular arrangements present in TPS.
- Due to the hydrophilic nature of potato fibers (presence of superficial hydroxyl groups) a
physical interaction with TPS is expected. We hypothesize that a potato fiber
incorporation will result in a modification of the intermolecular association present in
TPS. In conjunction, the dispersibility of TPS within the polyester phase could be altered
as well.
- We hypothesize that water uptake will be accelerated at higher potato fiber contents. This
effect is expected due to hydrophilic functionalities present in the potato fiber, as well as
the higher porosity within the material, which is supposed to reduce the barrier function
against water.

2. Experimental Section

2.1. Materials and Methods

TPS was prepared from corn starch (Agrana Stärke, Austria; raw material data: moisture =
max. 13 wt.% (weight percent), ash (550°C) = ~ 0.2 wt.% in dry substance, gelatinization
temperature = 80-85°C, amylose content= 27 wt.%, Figure S1), using glycerol (13 wt.%) as

2
plasticizer (purity 99 %; Brenntag, Austria). Liquid-dosing was conducted via a progressing
cavity pump (Netzsch, Germany). Experiments were carried out on a co-rotating twin screw
extruder (27D, Theysohn, Switzerland), using a screw speed of 200 revolutions per minute
(rpm), with a throughput of 10 kg per hour, and a temperature profile of
100/100/120/130/135/100/90 (from temperature zone 1 (feed section) to temperature zone 7
(die section)). The used screw configuration was equipped with alternating
shearing/conveying and kneading elements in order to ensure a sufficient plastification
(Figure S2). We added 1, 3 and 5 wt.% potato pulp residue derived fibers (produced by
Agrana Stärke, Austria), with these potato fibers having widths between 30-60µm, and
lengths (d50) between 100-200 µm, equaling in an aspect ratio of 4 ± 0.5, along with a
moisture content of 10 wt.% (Figure 1).

Figure 1: Scanning electron microscopy image of the used potato fiber (the spherical structures represent residual
starch granules)

Via hot-cutting uniformly shaped granules were obtained, which were subsequently re-
extruded with PBAT (BASF, Germany), and with 0.15 wt.% citric acid (Sigma Aldrich) as a
destructurization-agent (and potential compatibilizer), by applying the following processing
parameters: a mixing ratio of 1:1; screw speed 200 rpm, throughput 10 kg per hour, and a
temperature profile of 100/100/120/120/120/100/90 (from temperature zone 1 (feed section)
to temperature zone 7 (die section)). To accomplish a sufficient homogenization the screw
used for compounding was equipped with additional kneading elements (Figure S3). Powder
products (TPS-preparation) as well as polymer-granules (compounding) were fed into the
extrusion apparatus via pre-calibrated gravimetric dosing systems (Brabender, Germany).
With the obtained compounded material, flat films were produced on an OCS (Optical control
systems, Germany) film extrusion line (160 °C, 20 rpm, haul-off speed: 6.5 m/min). A film
thickness of 15-25 µm was obtained. TPS, compounds and film samples (Table 1) were all
analyzed for potato-fiber addition effects on material characteristics, at each processing step.

3
Table 1: Sample identification for the investigated potato-fiber contents, sample type, and control

Potato-Fiber (wt.%) TPS Sample Compound Sample Film Sample


0 Control
1 T1PF C1PF F1PF
3 T2PF C2PF F2PF
5 T3PF C3PF F3PF

2.2. TPS Characterization

Prior to characterization the samples were conditioned at 25 °C and 50 % relative humidity,


for 14 days.

2.2.1. Fourier transform infrared spectroscopy (FTIR)

Infrared spectra were recorded using a FTIR spectrometer (Bruker®, Alpha Sample
Compartment RT-DLaTGS, Austria) in attenuated total reflectance (ATR) mode, at a spectral
resolution of 2 cm-1, with the wavenumber range between 400 cm-1 and 4000 cm-1. The
presented FTIR spectra were each averaged from three independent measurements, which
were further vector normalized and baseline-corrected (rubberband method), using Bruker
software (OPUS version 7.5).

2.2.2. Viscoelastic properties

To obtain uniformly shaped rectangular samples for dynamic-mechanical analysis (DMA),


TPS was pressed (press type: Stroud-Glass, T.H. & J. Daniels). For this purpose, the TPS-
granules were heated to 150°C, and held at constant pressure of 8 bar (3 minutes), followed
by a 10 minute cooling cycle. The pressed samples were cut to a size of 10 mm x 60 mm, at a
thickness of 3.5 mm, and then conditioned at 25 °C and 50 % relative humidity.
Measurements were conducted on a Netzsch® DMA 242 C in 3-point bending mode, between
20°C to 160 °C, and at a heating rate of 10 °C / minute. Measurements were done in triplicate.

Storage modulus (E’), loss modulus (E’’) and loss factor (tan δ = ) were recorded. The

determination of relaxation processes associated with the glass transition were monitored via
the peak maxima of the loss modulus and tan δ curves. The required transition energy was

4
determined via the frequency sweep method (1; 5; and 10 Hz). A shift in glass transition
temperature (Tg) towards higher temperatures can be deduced from a demobilization of
polymeric molecules due to dynamic stress. A linear relationship is constituted, with ln (f,
frequency) and the reciprocal glass transition temperature ( ), (R represents the gas constant)

corresponding to the activation energy E a (equ.1) [15, 16]:

ln(f) = (equ.1)

2.2.3. X-ray diffractometry (XRD)

A Bruker® AXS GADDS D8 XRD-diffractometer with a 7-axis-goniometer (2θ, , , , and


x, y, z) and monochromatic X-ray radiation CuKα (λ=0.1542 nm) was used for the
determination of the X-ray diffraction pattern. The TPS granules were measured in reflection
mode ( = 12,5 °,  = 90 °,  = 0 °; 2θ = 12-24 °). Data analyses were done quantitatively, via
integration over the Debye Scherrer rings over angle 2θ.

2.3. Compound and film characterization

Samples were again conditioned at 25 °C, and 50 % relative humidity, for 14 days. Proper
sample preparation was applied for subsequent size exclusion chromatography, where the
compounded material was additionally extracted prior to the analytical characterization in
order to divide starch and polyester phases.

2.3.1. Size exclusion chromatography (SEC)

Weight average molecular weight (Mw), number average molecular weight (Mn) and
polydispersity (PDI) were determined with a Thermo Fisher Scientific chromatograph
(Dionex Ultimate 3000, Austria). Mw was used for comparative purposes, indicating
molecular weight related changes in a polymers mechanical properties [17]. For the
characterization of the polyester-phase the chromatograph was equipped with a styrene-
divinylbenzene copolymer column system (1000Å-1000000Å, chloroform, 0.6 ml/min), and a
refractive index detector (Refractomax 520), with polystyrene standards used for calibration
purposes. The determination of the TPS-phase was conducted with a
polyhydroxymethacrylate copolymer column system (100 Å -30000 Å, sodium nitrate,

5
0.7 ml/min). For sample detection a refractive index detector (RI-101) was used.
Monodisperse pullulan standards were applied for calibration purposes.

2.3.2. Light Microscopy/Scanning electron microscopy (SEM)

The compounds were fractured in liquid nitrogen and subsequently treated with 0.1 M
hydrochloric acid to remove the dispersed TPS phase [18, 19]. Prior to microscopic analysis
the samples were gold-coated using a sputter coater (JEOL®, JCM-1200 Fine Coater). The
particle surface was then examined in an electron microscope (JEOL®, JCM 5000 NeoScope,
10 kV, high-vacuum mode, China), which was equipped with a secondary electron detector.
The corn starch (raw material) as well as the TPS (monitoring of the obtained plastification
via polarization filter) were investigated by means of light microscopy (Olympus CX41RF,
Japan).

2.3.3. Differential scanning calorimetry (DSC)

The evaluation of the thermal compound characteristics happened via differential scanning
calorimetry (Mettler Toledo® 822, Austria). Measurements were conducted under constant
nitrogen flow of 50 ml/min, and a heating rate of 20 K/min. The first temperature cycle
involved a heating up to 200 °C, with a subsequent cooling to 20 °C. In a second cycle,
heating was applied up to 200 °C, followed by a cooling phase to -40 °C. The measurements
were run in triplicate. The crystallinity of the PBAT fraction in the compounded material was
determined with equation 2 (second heating cycle) [20, 21]:

100 (%) (equ.2)

where ΔHm is the enthalpy of fusion of the sample, is the heat of fusion for 100%
crystalline PBAT (114 J/g) [22], and f is the PBAT fraction. The accuracy of the added
PBAT-weight fraction (50 wt.%) was ensured via the utilization of pre-calibrated, gravimetric
dosing units. Furthermore, the concentration of the added PBAT fraction was roughly
monitored via FTIR spectroscopy (standard: pure PBAT; sample preparation: cryogenic
milling) (Figure S4).

2.3.4. Mechanical properties

For the determination of the mechanical film properties a Zwick Roell (Germany) universal
testing machine was used. Measurement and sample geometry followed ÖNORM EN ISO
527-3 standard (clamp distance 100 mm). The tensile modulus (secant modulus, slope of the
6
elastic region in the stress-strain curve) was tested under a cross-head speed of 1 mm/min
(with extensometer). A cross-head speed of 100 mm/min was applied for the measurement of
tensile strength and elongation at break. For each film sample at least ten specimens were
tested (in machine direction = longitudinal production direction).

2.3.5. Dynamic vapor sorption analysis

Film material sorption/desorption characteristics were determined with the dynamic vapor
sorption apparatus DVS Advantage® (Surface Measurement Systems Ltd., UK). A thin film
was suspended in the DVS instrument and the sorption kinetics for a series of humidity steps
were recorded. After initial drying, the sample was gradually exposed to increasing steps in
humidity from 0 % relative humidity (RH) to 98 % RH and back to 0 % RH (temperature
23 °C).

2.3.6. Statistical evaluation

Analysis of variance ANOVA was conducted, followed by Tukey multiple comparison post-
hoc data analyses. The results were evaluated according to statistically significant changes
(significant difference at p < 0.05), using Minitab 17.0.

3. Results and discussion

3.1. TPS Characterization

3.1.1 Structural characterization

Figure 2 shows the FT-IR spectra in the wavenumber range between 400 and 4000 cm-1. The
most significant change with respect to the addition of potato fibers is a peak intensity
increase seen at 3300 cm-1, which is an indication of an enhanced hydrogen-bonds
appearance. The signals dedicated to the ether linkage at 993 cm-1 increased with fiber
content, which indicates a stronger intensity of the C-O-C stretching vibration along with
fiber contents. This rise in intensity could be attributed to altered intramolecular
configurations, respectively, to changes in molecular packing [23-25].

7
Figure 2: FTIR spectra of potato-fiber modified TPS samples, compared to the control (no fibers added)

3.1.2 Viscoelastic properties

Table 2 shows the storage modulus (E’), the loss modulus (E’’), the tan δ, and the phase
transition activation energy (Ea) of the potato fiber modified TPS samples, compared with the
control.
Table 2: Viscoelastic properties of potato fiber modified TPS (T1PF, T2PF, T3PF = 1, 3, and 5 wt.% potato
fibers added) and the control (no fibers added)

Maximum
Storage Modulus E’ (MPa) at specified Maximum of Loss Activation
of Loss
temperature-points Modulus E’’ Energy
Factor tan δ
20°C 60°C 100°C °C MPa °C kJ/mol
a a a a a a
Control 2014.1 ± 20.1 1239.0 ± 23.4 62.4 ± 1.8 67.9 ± 2.3 170.0 ± 15.6 88.3 ± 0.4 432.8 ± 14.7b

T1PF 1958.8 ± 155.1a 1592.0 ± 162.9ab 325.4 ± 53.db 82.3 ± 2.7b 172.0 ± 34.0a 109.1 ± 5.0b (351.9 ± 32.5a)

T2PF 1749.6 ± 233.8a 1987.1 ± 123.7b 807.4 ± 89.0c 91.1 ± 4.0bc 182.0 ± 7.1a 130.8 ± 1.9c 454.8 ± 23.1b

T3PF 1898.4 ± 285.7a 2018.3 ± 109.9b 1116.5 ± 87.3d 99.4 ± 3.9c 250.0 ± 15.4b 144.0 ± 3.3d 541.2 ± 11.9c

Numbers followed by the same letter are not significantly different (significant difference at p < 0.05); See Table
1 for sample identification

Data show that the incorporation of potato fibers led to higher storage moduli (60, 100 °C),
and to higher loss moduli. This effect can be paralleled with the formation of a tighter
molecular network along with increased stiffness. In general, the measurement curves show
features that are characteristic for partially miscible TPS-glycerol systems (e.g. temperature
shift between loss modulus/loss factor maxima) [26]. With increasing temperature the free
volume is expected to increase, whereas in the case of the control, T1PF and T2PF only one
8
distinct tan δ peak was visible (Figure S5). Thus, it can be concluded that large chain-
segments within the amorphous regions in TPS directly started to move (multi-chain, large-
scale motions), which is referred to as α-transition or Tg. In case of the sample T3PF the
measurement curves exhibited a tan δ peak-shoulder. The mechanism can be attributed to a
minor transition effect that overlaps with/directly passes into the main glass transition peak
[27]. Generally, it has to be taken into account that tan δ depends upon various factors, like
for example: matrix/fiber interaction, fiber flexibility and fiber/fiber contact [28]. Compared
to the control, no significant changes in tan δ peak height were observed at the different fiber
concentrations. An increase of the tan δ peak temperature at higher fiber contents implies
increases in Tg and consequently also in stiffness. This is evidence for a fiber-induced
reinforcing effect, and also for the existence of a relationship between fiber content and the
activation energy. With regard to the structural characteristics of TPS the observed effect is
attributed to a reduced flexibility of the main molecule-chains, respectively, to a reduction in
free volume (chain-immobilization in the presence of the fiber) [29-31]. Furthermore an
additional mechanism could have contributed to the observed effect: At the higher fiber
concentrations hydrogen bonding may have contributed to the formation of less flexible
linkages between the starch matrix and the potato fibers. As a consequence, the interaction
between TPS and the plasticizer glycerol became disturbed, which in turn affected the
materials flexibility. At a fiber concentration of 1 wt.% the activation energy exhibited a
slight decrease, which can be deduced from an inhomogeneous distribution of the potato
fibers in the TPS matrix. On closer examination, the ln(frequency)/inverse temperature plot in
case of T1PF did not fulfill the requirement of a coordinated relationship (vague linearity in
comparison with control, T2PF and T3PF). The measurement was repeated, however, the
inconsistent trend remained. Thus, the calculation for T1PF was not taken into account in the
final activation-energy evaluation (result in brackets, Table 2). Sample T3PF exhibited a
remarkable rise in activation-energy, which can be deduced from an increase in material-
stiffness. This implies that, due to interactions between fiber and matrix, more energy is
required to induce molecular movement [32].

9
Graphical analysis of the observed changes
(Activation Energy plots, x-axis shifted for Control T1PF
comparative purposes)

Control T2PF T3PF

2.5

1.5 ccc Vague linear relationship


ln(f)

T2PF T3PF
1

0.5

0
1/T

Figure 3: Activation energy determination following the frequency dependency (1; 5; 10 Hz) of tan δ,
R2=coefficient of determination; See Table 1 for sample identification

3.1.3 Molecular weight distribution

Table 3 summarizes the weight average molecular weights of the fiber-modified TPS
material, and the control.
Table 3: Relative molecular weight (weight average molecular weight Mw) of potato fiber modified TPS in
comparison to the unmodified control (recovery rate; 78, 83, 80, 84 % – Control  T3PF)

Mw (kDa)
Control 1560.0 ± 5.0a
T1PF 1270.0 ± 1.5b
T2PF 1290.0 ± 8.5b
T3PF 1230.0 ± 10.1b
Numbers followed by the same letter are not significantly different (significant difference at p < 0.05); See Table
1 for sample identification

It is evident that the alignment of the potato fibers during extrusion has altered the rheological
behavior of the melt, which furthermore led to an increase in shear stresses. The mechanism
has been also reported in the literature [33]. An increase in specific mechanical energy was

10
observed, which resulted in a shift of the molecular weight distribution towards lower values
(control  fiber-modified TPS). This effect remained at the same level, even at higher fiber
contents (Figure S6).

3.1.4 Molecular arrangement

Figure 4 comparatively shows the XRD measurements of unmodified TPS, and of potato fiber
modified TPS. The patterns exhibit two well-defined peaks: a distinct peak at a diffraction
angle of 19.3 °, and a smaller one at 12.6 °. As stated in the literature [34] there are two
characteristic diffraction peaks in the case of glycerol-plasticized TPS, which refer to the so
called V-type crystalline structure. In this study, the inherent crystallinity of the starch granule
was destroyed during extrusion (see Figure S7), so the occurrence of residual crystallinity due
to intact granules [35] was excluded. Amylose has the ability to form process-induced
crystallinity, in conjunction with small additive molecules, i.e. plasticizers such as glycerol, or
other processing aids [36, 37]. The characteristic appearance of peaks at certain XRD angles
is used as a criterion to distinguish between the non-hydrated (VA), the hydrated (VH) and the
intermediate (EH) crystal form [35]. The visibility of peaks at angles around 13 ° and 20 ° was
assigned to the VH type crystalline structure [36]. The peak at a diffraction angle of 12.6 ° was
weakened at fiber concentrations of 1 wt.% and 3 wt.%, and finally disappeared at a
concentration of 5 wt.%. Literature [37] reported, that the shielding of polar groups due to
derivatization inhibits amylose recrystallization. A further reference [38] described the
inhibition of regular molecular packing by means of well-dispersed filler particles. A reason
for the disappearance of the peak at an angle of 12.6 ° could be a disturbance in crystal
formation, respectively, a reduced helix formation because of the potato fibers, which
probably act like steric hindrances.

11
Figure 4: X-Ray diffraction patterns of potato fiber modified TPS and the control ((T1PF, T2PF, T3PF = 1, 3,
and 5 wt.% potato fibers added)

3.2. Compound and film characterization

3.2.1 Molecular weight distribution

It is evident that potato fiber addition has lowered the average molecular weight of the starch
fraction, while the polyester fraction remained unaffected (Table 4). This allowed the
conclusion that the potato fiber accelerated the disintegration of starch, as it has mitigated the
protective shielding function of the polyester-melt, which made starch more susceptible to
shear degradation [39]. The mechanism describes the reduction of the molecular weight in
starch during extrusion via partial debranching and shear scission [40]. It is evident that the
degradation-mechanism mainly affected the high-molecular weight fraction in TPS (Figure
S8).
Table 4: Relative molecular weights (weight average molecular weight Mw) of the two compound constituents;
1) potato fiber modified starch and the control, 2) polyester PBAT

Starch PBAT
Sample Mw (kDa)
a
Control 240.0 ± 7.0 129.5 ± 0 9a
C1PF 204.5 ± 23.5a 130.1 ± 4 0a
C2PF 117.5 ± 17.7b 119.2 ± 3 1a
C3PF 84.2 ± 26.6b 120.0 ± 3.5a
Numbers followed by the same letter are not significantly different (significant difference at p < 0.05); See Table
1 for sample identification

12
3.2.2 Phase transition behavior

The amorphous TPS phase exhibited two separate glass transitions (Table 5, Figure S9). This
is attributed to plasticizer-rich, and to plasticizer-poor domains prevalent in TPS, which
evolved due to the different arrangement/orientation of chain segments, followed by a phase
separation [26]. Overlay effects most likely restricted the visibility of the two transitions
obtained in the DMA analysis, as only one tan δ peak was visible. Another possibility is the
fact, that phase-separation could have occurred later on, during the course of compounding.
As the TPS-glass transition directly passes into PBAT-melting, it was not possible to
determine the PBAT-melting peak characteristics in the first temperature cycle. Hence, the
melting peak temperature (Tm) along with the melting peak onset temperature (Tm ONSET) were
determined both via the thermograms obtained in the second temperature cycle (overlay
effects due to TPS widely eliminated) (Figure S9).
Table 5: Phase transitions (Tg – glass transition temperature, Δcp – change in heat capacity, Tm- melting
temperature, Hm – melting enthalpy, Xc – crystallinity) observed in compound materials based on the control and
potato fiber modified TPS

TPS PBAT
st
1 cycle 2nd cycle
Sample Tg1 (°C) Δcp1 (J/gK) Tg2 (°C) Δcp2 (J/gK) Tm ONSET (°C) Tm (°C) Hm (J/g) Xc (%)
a a a a a
Control - - 76.5 ± 1.5 0.213 ± 0.002 94.6 ± 0.8 119.5 ± 0.5 7.2 ± 0.2 12.6 ± 0.2a

C1PF 42.7 ± 0.6a 0.161 ± 0.013a 76.6 ± 0.7a 0.258 ± 0.015b 94.3 ± 0.3a 121.5 ± 0.6b 6.3 ± 0.1b 10.9 ± 0.0b
C2PF 42.6 ± 1.0a 0.158 ± 0.014a 77.0 ± 0.7a 0.244 ± 0.018b 93.1 ± 0.9a 120.4 ± 0.1ab 6.4 ± 0.1b 11.2 ± 0.1b
C3PF 42.4 ± 1.1a 0.175 ± 0.012a 76.7 ± 0.8a 0.257 ± 0.006b 92.5 ± 0.9a 119.9 ± 1.0a 6.9 ± 0.2a 12.1 ± 0.3a

Numbers followed by the same letter are not significantly different (significant difference at p < 0.05); See Table
1 for sample identification
Apart from the overall crystallinity, the melting process is influenced by the crystallite size
and the crystallite size distribution. Tm and Tm ONSET showed only minor differences (second
cycle). The crystallinity calculation for PBAT was performed with the data obtained from the
second heating cycle. Only small, negligible changes in crystallinity were observed. The peak
shape in the second heating cycle was similar for the investigated samples (no indication for
changes in crystallite size distribution). Δcp was taken as an indicator for the structural
arrangement within TPS, as the heating capacity step directly relates to the presence of
disordered, amorphous regions. The observed increase in Δcp (comparison control versus
fiber-modified samples) leads to the conclusion that the fiber-addition has promoted the
growth of disordered TPS-structures [41-46].

13
3.2.3 Microscopic compound structure/Film surface

Figure 5 shows etched fracture surfaces of the potato fiber modified materials, and the
control, visualizing the distribution of the TPS inside of the PBAT-matrix.

Control-Compound C1PF C2PF C3PF

Control-Film F1PF F2PF F3PF

100 µm

Figure 5: Fractured TPS-polyester compounds etched with hydrochloric acid. Distribution of the dispersed TPS
phases (SEM), and the relationship between fiber content and compound- versus film-homogeneity (reflected-
light microscopy); See Table 1 for sample identification

The etched control shows TPS-cavities that are distributed as small, regular particles scattered
throughout the PBAT-matrix. The fine dispersion of TPS within the PBAT-matrix can be
attributed to the complete destructurization of the starch granules (Figure S7), the
plastification (reduction in intermolecular friction, inhibition of hydrogen-bonding, increase in
molecule mobility) with glycerol, and the therefrom achievable improved flowability of the
TPS, which facilitates the miscibility of the two phases. The pictures indicate that the potato
fibers affected the miscibility and thus, induced agglomeration and the formation of irregular
particles (increased particle size along with increasing fiber content). This agglomeration
impeded the otherwise homogeneous and fine distribution of the dispersed phase within the
compounded material and, at elevated fiber concentration (F3PF), affected the microstructure
of the produced films.

3.2.4 Mechanical properties

Tensile properties of the flat films are compiled in Figure 6.

14
500 25
Elongation (%) / Modulus (MPa)

a b

Tensile Strength (MPa)


400 20
a c
300 b 15
a b,c
c
200 10

a b
100 c 5
d
0 0
Control F1PF F2PF F3PF
Tensile Modulus (MPa)
Elongation at Break (%)
Tensile Strength (MPa)

Figure 6: Tensile properties of control and potato fiber modified films; values labelled with the same letter are
not significantly different (significant difference at p < 0.05); See Table 1 for sample identification

The incorporation of the fiber obviously provoked an increase in tensile modulus, which can
be attributed to the rigid nature of the fiber and the associated reduction in chain mobility. As
the properties of the compound with increasing fiber content came close to those of neat TPS,
i.e. having reduced elongation and increased stiffness, it is also likely that the potato fibers
improved the amylose and amylopectin incorporation within the polyester matrix [47]. On the
other hand, the potato fiber incorporation in TPS has led to reduced film-elongation and, at
elevated fiber content (5 %), to a reduction in tensile strength. Literature reports that fillers
may agglomerate and thus, form (or promote the formation of) defects (the effect usually is
enhanced at elevated fiber content) [48]. In addition, a reduction in molecular weight could
have affected the mechanical characteristics [49]. As the molecular weight of the starch
fraction in TPS was reduced, this might have contributed to the observed changes in
mechanical behavior [10, 50, 51].

3.2.5 Sorption characteristics

Data on the influence of potato fiber contents on sorption characteristics are shown in Figure
7.

15
Control F1PF F2PF F3PF
45 45 45 45

40 40 40 40

35 35 35 35
Mass Change (%)

30 30 30 30

25 25 25 25

20 20 20 20
15 15 15 15
10 10 10 10
5 5 5 5
0 0 0 0
0 50 100 0 50 100 0 50 100 0 50 100
RH (%) RH (%) RH (%) RH (%)

Figure 7: Sorption isotherms of the potato fiber modified TPS/polyester films (─ Desorption; --- Sorption), and
the control; See Table 1 for sample identification

It can be seen that the potato fiber additions increased the equilibrium water content at full
RH saturation (control: 37 % mass change at 98 % RH; F3PF: 42 % mass change at 98 %
RH). Furthermore, the potato fibers seemingly triggered a hysteresis between the sorption and
the desorption isotherms. It can be concluded that the potato fiber additions have created an
amorphous, hydrophilic TPS-network, which in turn fostered the water sorption as well as
water retention characteristics of the tested potato fiber – modified TPS/polyester films [52,
53].

4. Conclusions

The conducted research on the TPS material revealed that the addition of potato fibers
significantly influenced the structural arrangement. This became also evident with recorded
mechanical data (DMA analysis), which exhibited prevalent rigidity when potato fibers were
added. The effect can be attributed to the formation of less flexible linkages between starch
and fiber, respectively, to a reduction in free volume (chain-immobilization in the presence of
the fiber). The modification had a decisive influence on the processing-induced molecular
arrangement of amylose and amylopectin, while XRD analysis demonstrated that the
thorough formation of crystalline areas was also affected by the added potato fibers.

The characterization of the compound revealed that the addition of potato fibers has led to
TPS-fiber agglomerations. This has also affected the fine distribution of the dispersed phase
within the compound and furthermore, the microstructure of the produced film materials. SEC
results indicated that the fiber additive induced additional shear, which led to enhanced TPS
disintegration, while the polyester phase remained unaffected (hypothesis 1). DSC analysis
has shown that the TPS phase in the compounded material was present in two different stages,
(1) a plasticizer-rich, and (2) a plasticizer-poor one. This phenomenon has resulted in two
16
glass transitions. The observed change in heat capacity led to the conclusion that the fiber-
addition has promoted the growth of disordered TPS-structures. The transition characteristics
of the polyester phase exhibited only negligible changes in crystallinity.

The mechanical film properties showed enhanced stiffness at reduced elongation, both with
increasing fiber content. This observation can be deduced from a combined effect, which
involves: 1) an increase in tensile modulus due to the rigid nature of the fiber and the
associated reduction in molecule-mobility, as well as the good incorporation of amylose and
amylopectin within the polyester matrix; 2) a reduction in elongation due to the formation of
agglomerates that acted like defects, respectively, promoted the formation of defects. A
reduction in TPS-molecular weight, might have contributed to the observed changes in
mechanical behavior (hypothesis 2). The issue of agglomeration could potentially be solved
by adding appropriate compatibilizing-agents, which are capable of improving the separation
of the fiber material and furthermore enable a homogeneous distribution within the matrix.
The utilization of potato fibers in TPS resulted in the formation of an amorphous, hydrophilic
TPS-network, which in turn facilitated film water uptake (hypothesis 3).

Table 6: Summary of the demonstrated changes in TPS, compound and films, along with increasing potato fiber
contents (< - reduced, > - increased, = - remained at the same level)

TPS Compound Film


TPS PBAT
Hydrogen bonding (Tg) > = - -
Crystallinity < < = -
Molecular weight < < = -
Stiffness > - - >
Hydrophilicity - - - >
The obtained results (summary Table 6) are crucial in follow-up research, especially in
optimizing sorption characteristics, strength and haptic properties of TPS-based film
materials, with the goal of moving towards functional packaging applications.

Acknowledgement

The authors would like to thank the Austrian Research Promotion Agency (FFG, Project
Number: 854577) for the financial support. Furthermore, the authors wish to thank Agrana
Stärke for generously supplying the required starch and potato fiber raw materials.

17
Conflict of Interest

The authors declare no conflict of interest.

The raw/processed data required to reproduce these findings cannot be shared at this time due
to technical or time limitations.

References

1. Brandelero, R.P., M.V. Grossmann, and F. Yamashita, Films of starch and


poly(butylene adipate co-terephthalate) added of soybean oil (SO) and Tween 80.
Carbohydr Polym, 2012. 90(4): p. 1452-60.
2. Ren, J., et al., Preparation, characterization and properties of binary and ternary
blends with thermoplastic starch, poly(lactic acid) and poly(butylene adipate-co-
terephthalate). Carbohydr Polym, 2009. 77(3): p. 576-582.
3. Hablot, E., et al., Reactive extrusion of glycerylated starch and starch–polyester graft
copolymers. Eur Polym J, 2013. 49(4): p. 873-881.
4. Shirai, M.A., et al., Thermoplastic starch/polyester films: effects of extrusion process
and poly (lactic acid) addition. Mater Sci Eng C. Mater Biol Appl, 2013. 33(7): p.
4112-7.
5. Olivato, J.B., et al., Effect of organic acids as additives on the performance of
thermoplastic starch/polyester blown films. Carbohydr Polym, 2012. 90(1): p. 159-64.
6. Lin, C.-A. and C.-C. Tung, The Preparation of Glycerol Pseudo-Thermoplastic Starch
(GTPS) via Gelatinization and Plasticization. Polym Plast Technol Eng, 2009. 48(5):
p. 509-515.
7. Niazi, M.B.K., M. Zijlstra, and A.A. Broekhuis, Influence of plasticizer with different
functional groups on thermoplastic starch. J Appl Polym Sci, 2015. 132(22): p. n/a-
n/a.
8. Edhirej, A., et al., Effect of various plasticizers and concentration on the physical,
thermal, mechanical, and structural properties of cassava-starch-based films. Starch -
Stärke, 2017. 69(1-2): p. 1500366.
9. Pérez-Pacheco, E., et al., Thermoplastic Starch (TPS)‐ Cellulosic Fibers Composites:
Mechanical Properties and Water Vapor Barrier: A Review. 2016, INTECH.
10. Ma, X., J. Yu, and J.F. Kennedy, Studies on the properties of natural fibers-reinforced
thermoplastic starch composites. Carbohyd Polym, 2005. 62(1): p. 19-24.
11. Rommi, K., et al., Potato peeling costreams as raw materials for biopolymer film
preparation. J Appl Polym Sci, 2016. 133(5): p. n/a-n/a.
12. Averous, L.U.C., & Fringant, C., Association Between Plasticized Starch and
Polyesters: Processing and Performances of Injected Biodegradable Systems. Polym
Eng Sci, 2001. 41(5): p. 727-734.
13. Mei, X., T.H. Mu, and J.J. Han, Composition and physicochemical properties of
dietary fiber extracted from residues of 10 varieties of sweet potato by a sieving
method. J Agric Food Chem, 2010. 58(12): p. 7305-10.
14. Pastuszewska, B., et al., POTATO DIETARY FIBRE – PRELIMINARY
CHARACTERIZATION OF THE PROPERTIES AND NUTRITIONAL EFFECTS – A

18
REVIEW. POLISH JOURNAL OF FOOD AND NUTRITION SCIENCES, 2009.
59(3): p. 205-210.
15. Siengchin, S., Thermomechanical Analysis and Processing of Polymer Blends, in
Characterization of Polymer Blends: Miscibility, Morphology and Interfaces. 2015,
Wiley-VCH Verlag GmbH & Co KGaA: Weinheim. p. 393-414.
16. Fahrngruber, B., et al., Malic acid: A novel processing aid for thermoplastic
starch/poly(butylene adipate-co-terephthalate) compounding and blown film
extrusion. J Appl Polym Sci, 2017. 134(48).
17. Totten, G.E., Fuels and Lubricants Handbook, ed. S.R.W. Rajesh J. Shah, George E.
Totten. 2003, West Conshohocken: ASTM International.
18. Ali Nezamzadeh, S., Z. Ahmadi, and F. Afshari Taromi, From microstructure to
mechanical properties of compatibilized polylactide/thermoplastic starch blends. J
Appl Polym Sci, 2017. 134(16).
19. Ma, P., et al., Tailoring the morphology and properties of poly(lactic
acid)/poly(ethylene)-co-(vinyl acetate)/starch blends via reactive compatibilization.
Polym Int, 2012. 61(8): p. 1284-1293.
20. Donovan, J.W., Phase Transitions of the Starch-Water System. Biopolymers, 1979.
18: p. 263-275.
21. Vertuccio, L., et al., Nano clay reinforced PCL/starch blends obtained by high energy
ball milling. Carbohydr Polym, 2009. 75(1): p. 172-179.
22. Al-Itry, R., K. Lamnawar, and A. Maazouz, Improvement of thermal stability,
rheological and mechanical properties of PLA, PBAT and their blends by reactive
extrusion with functionalized epoxy. Polym Degrad Stab, 2012. 97(10): p. 1898-1914.
23. Pavia, D.L., et al., Introduction to Spectroscopy. 4 ed. 2009, Belmont: Brooks/Cole
Cengage Learning.
24. Stuart, B., Infrared Spectroscopy: Fundamentals and Applications. 2004, West
Sussex: John Wiley & Sons Ltd.
25. Synytsya, A. and M. Novak, Structural analysis of glucans. Ann Transl Med, 2014.
2(2): p. 17.
26. Castillo, L., et al., Thermoplastic starch films reinforced with talc nanoparticles.
Carbohydr Polym, 2013. 95(2): p. 664-74.
27. Battegazzore, D., et al., Isosorbide, a green plasticizer for thermoplastic starch that
does not retrogradate. Carbohydr Polym, 2015. 119: p. 78-84.
28. Sreenivasan, V.S., et al., Dynamic mechanical and thermo-gravimetric analysis of
Sansevieria cylindrica/polyester composite: Effect of fiber length, fiber loading and
chemical treatment. Compos Part B-Eng, 2015. 69: p. 76-86.
29. Mohanty, S. and S.K. Nayak, Biodegradable nanocomposites of poly (butylene
adipate-co-terephthalate) (PBAT) with organically modified nanoclays. Int J Plast
Technol, 2011. 14(2): p. 192-212.
30. Wang, N., J. Yu, and X. Ma, Preparation and characterization of thermoplastic
starch/PLA blends by one-step reactive extrusion. Polym Int, 2007. 56(11): p. 1440-
1447.
31. Ward, I.M.S., J., An Introduction to the Mechanical Properties of Solid Polymers.
2004, England: John Wiley & Sons Ltd.
32. Khouloud, J.M., M.C.; Sabu, T., Clay-Polymer Nanocomposites. 2017, United States:
Elsevier.
33. Carvalho, A.J.F.Z., M.D.; Curvelo, A.A.S; Gandini, A., Size exclusion
chromatography characterization of thermoplastic starch composites 1.Influence of
plasticizer and fibre content. Polym Degrad Stab, 2003. 79: p. 133-138.

19
34. Shi, R., et al., Characterization of citric acid/glycerol co-plasticized thermoplastic
starch prepared by melt blending. Carbohydr Polym, 2007. 69(4): p. 748-755.
35. van Soest, J.J.G., Hulleman, S.H.D., de Wit, D., Fliegenthart, J.F.G., Crystallinity in
Starch Bioplastics. Ind Crops Prod, 1996. 5: p. 11-22.
36. Ren, J., et al., Characteristics of starch-based films produced using glycerol and 1-
butyl-3-methylimidazolium chloride as combined plasticizers. Starch - Stärke, 2017.
69(1-2): p. 1600161.
37. Chaudhary, A.L., et al., Amylose content and chemical modification effects on
thermoplastic starch from maize – Processing and characterisation using
conventional polymer equipment. Carbohydr Polym, 2009. 78(4): p. 917-925.
38. Magalhães, N.F. and C.T. Andrade, Calcium Bentonite as Reinforcing Nanofiller for
Thermoplastic Starch. J Braz Chem Soc, 2010. 21(2): p. 202-208.
39. R. Chinnaswamy and M.A. Hanna, PHYSICOCHEMICAL ANO
MACROMOLECULAR PROPERTIES OF STARCH-CELLULOSE FIBER
EXTRUDATES. FOOD STRUCTURE, 1991. 10: p. 229-239.
40. Liu, W.-C., P.J. Halley, and R.G. Gilbert, Mechanism of Degradation of Starch, a
Highly Branched Polymer, during Extrusion. Macromolecules, 2010. 43(6): p. 2855-
2864.
41. Tajuddin, S., et al., Rheological properties of thermoplastic starch studied by
multipass rheometer. Carbohydr Polym, 2011. 83(2): p. 914-919.
42. Ehrenstein, G.W., G. Riedel, and P. Trawiel, Thermal Analysis of Plastics - Theory
and Practice. 2004, München: Carl Hanser Verlag.
43. Mano, J.F., D. Koniarova, and R.L. Reos, Thermal properties of thermoplastic starch/
synthetic polymer blends with potential biomedical applicability. J Mater Sci Mater
Med, 2003. 14: p. 127-135.
44. C.G. Biliaderis*, A.L., I. Arvanitoyannis, Glass transition and physical properties of
polyol-plasticised pullulan–starch blends at low moisture. Carbohydr Polym, 1999.
40: p. 29-47.
45. K SREENIVASAN, P.D.N.a.V.V.B., Differential scanning calorimetric studies of
polyester fabrics used in sewing ring of an heart valve. Bull Mater Sci, 1983. 5(2): p.
123-126.
46. Da Róz, A.L., et al., Thermoplastic starch modified during melt processing with
organic acids: The effect of molar mass on thermal and mechanical properties. Ind
Crop Prod, 2011. 33(1): p. 152-157.
47. Jawaid, M.S., S.K., Bionanocomposites for Packaging Applications. 2018,
Switzerland: Springer International Publishing.
48. Poletto, M., Polystyrene cellulose fiber composites: effect of the processing conditions
on mechanical and dynamic mechanical properties. Matéria (Rio de Janeiro), 2016.
21(3): p. 552-559.
49. Chabrat, E., et al., Influence of citric acid and water on thermoplastic wheat
flour/poly(lactic acid) blends. I: Thermal, mechanical and morphological properties.
Ind Crop Prod, 2012. 37(1): p. 238-246.
50. Rhim, J.-W., Potential use of biopolymer-based nanocomposite films in food
packaging applications. Food Sci Biotechnol, 2007. 16(5): p. 691-709.
51. Mittal, V., T. Akhtar, and N. Matsko, Mechanical, Thermal, Rheological and
Morphological Properties of Binary and Ternary Blends of PLA, TPS and PCL.
Macromol Mat Eng, 2015. 300(4): p. 423-435.
52. Enrione, J.I., S.E. Hill, and J.R. Mitchell, Sorption and Diffusional Studies of Extruded
Waxy Maize Starch-Glycerol Systems. Starch-Stärke, 2007. 59(1): p. 1-9.

20
53. Oliver, L. and M.B.J. Meinders, Dynamic water vapour sorption in gluten and starch
films. J Cereal Sci, 2011. 54(3): p. 409-416.

21
Highlights
Sustainable potato fiber, a biorefinery by-product, as green additive in biopolymers
Biobased and biodegradable additive for thermoplastic starch (TPS)
Overall characterization of TPS and fiber modified TPS (Microscopy, SEC, XRD, FTIR,
DMTA)
The fiber interacts with thermoplastic starch to form an amorphous, hydrophilic network
Structure-property relationship in TPS influences compound and film properties
Film material with haptic and mechanical properties suitable for industrial applications

22
3

Starch + Plasticizer + Potato Fiber 2,5


Structure-Property 2

Stress
Relationship 1,5
Tensile Modulus
1
0,5
0
0 1 Strain
2 3 4

Tensile Modulus Increase


Cellulose, Hemicellulose
Lignin, Pectin
Changes in Starch Crystallization
- Molecular Weight
Thermoplastic Starch (TPS)
Hydrophilic
Functionalities
Thermoplastic Starch + Polyester Potato
Fibers

TPS Degradation

Water-Uptake Increase
45

Moisture Content
40

35

30

25

20

15
Biopolymer Compound 10

0
0 20 40 60 80 100
Relative Humidity

You might also like