Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Journal of Polymers and the Environment

https://doi.org/10.1007/s10924-020-01874-w

ORIGINAL PAPER

Biodegradable Polymer Blends Based on Thermoplastic Starch


Vesna Ocelić Bulatović1   · Vilko Mandić2 · Dajana Kučić Grgić2 · Antonio Ivančić1

© Springer Science+Business Media, LLC, part of Springer Nature 2020

Abstract
This work considered the utilisation of environmentally friendly biodegradable materials instead of conventional polymeric
materials, in order to prevent further environmental endangerment by accumulation of non-biodegradable materials. Ther-
moplastic starch (TPS) is an exemplary environmentally friendly biodegradable material. Blending TPS with other biode-
gradable polymers such as biodegradable polyesters; polylactic acid (PLA) and polycaprolactone (PCL), was recognised as
a successful strategy to provide a renewable, fully biodegradable and cost-effective materials. Namely, PLA, PCL and TPS
polymeric materials individually show some shortcomings, which can be surpassed by blending. Using economically viable
procedure, one can produce multicomponent polymeric materials complying with functional performance requirements
while achieving complete biodegradation. Accordingly, PLA/PCL/TPS ternary blends with compositional variations were
prepared and investigated in terms of structural, morphologic, mechanical, thermal, physicochemical (water absorption,
contact angle) and processing properties in order to discuss and elucidate optimal performance of the overall material. In
summary, this study could lead to a better understanding of performance behind different blends and ensure producing and
using a new generation of biodegradable plastics that will not burden our environment.

Keywords  Biodegradable materials · Thermoplastic starch · Polylactic acid · Polycaprolactone

Introduction reducing carbon dioxide emissions [1, 2]. Biopolymers are


abundant and economically favourable. Despite the increas-
Nowadays, an increasing environmental awareness of the ing use of biopolymers and increased production capacity,
public and stakeholders finally allowed a shift of the focus the environmental impact on behalf of biopolymer advanta-
of the scientific community towards the development of bio- geous properties is still not fully utilised. Some drawbacks
degradable materials from renewable sources with the aim of biopolymers still hinder them to meet the market expecta-
of completely replacing conventional non-biodegradable tions [3]. Biopolymers are still quite expensive compared to
materials in many applications. Namely, the predominate conventional fossil fuel based polymers. Consequently, mod-
approach to use petroleum-based plastic packaging that end- ifying the biopolymer materials to yield the desired perfor-
up in landfills where recycling is difficult or restricted needs mance is currently the hot topic in the scientific community.
to be reduced. Recently, the plastics industry displayed an Unlike demanding development of new polymer materials
interest in finding alternative sources of raw materials. Many or polymerization pathways, blending is a relatively inex-
natural materials were considered [1], however polymers pensive and fast answer on how to adjust the plastic product
produced from natural renewable raw materials, i.e. biopol- properties. Namely, modification by means of blending of
ymers, turned out to be particularly interesting. Namely, selected bio-based and/or biodegradable materials offers the
biopolymers can successfully substitute for fossil sourced ability to adjust properties over a wide range while comply-
polymers and therefore can contribute to the environment by ing with low carbon footprint legislative constrains.
Poly (lactic acid) (PLA) is a thermoplastic material that
belongs to the group of biodegradable polyesters. PLA high-
* Vesna Ocelić Bulatović
vocelicbu@simet.hr
lights repose on its extreme stiffness, tensile strength and
gas permeability as well as other outstanding mechanical
1
University of Zagreb Faculty of Metallurgy, Sisak, Croatia properties in line with the properties of synthetic polymers,
2
University of Zagreb Faculty of Chemical Engineering perfectly allowing packaging applications. On the other
and Technology, Zagreb, Croatia

13
Vol.:(0123456789)
Journal of Polymers and the Environment

hand, the limitations of PLA are found in low degradation material. Thus, blending TPS with biodegradable polyesters
rate, relatively high cost and high inherent brittleness [1, may be an excellent way to provide a renewable, fully bio-
4], which prevent broader applications. Addressing the brit- degradable and cost-effective material.
tleness problem of PLA would broaden its applicability, Previous studies have shown that blending PCL and TPS
justifying the focus of recent investigations. Blending PLA increases PCL biodegradation rate. Low degradation rate of
with flexible polymers such as poly-ε-caprolactone (PCL), PLA may be increased by blending with other biodegradable
poly-buthylene succinate (PBS), and poly-butylene succi- polymers that have higher degradation rates. Natural starch
nate adipate (PBSA) was investigated to clarify the extent can be a good blending candidate for enhancing the biodeg-
of upgrades of flexibility and elongation strain at break [1, radability of PLA and PCL due to its low cost, abundance
5–8]. PCL was shown to be a good candidate to address the and availability from various botanical resources [19, 20].
shortcomings of PLA. Although the success of binary blends has been docu-
Poly-ε-caprolactone (PCL) is highly flexible biopolymer mented, recently, industrial and academic research is con-
with extremely high elongation at breakage and low ten- cerned with the study and development of multicomponent
sile strength [9]. PCL structure resembles combination of polymer blends consisting of three or more components [21].
ordered crystalline domains and non-ordered domains con- Namely, broader range of useful properties may be achieved
sequently allowing rigid-rubbery properties; such configura- if a successful balance of the constituents is provided. By
tion has been widely acknowledged for improving elasticity. combining different materials with complementary proper-
PCL is a hydrophobic semi-crystalline polymer with broad ties, multicomponent polymeric materials may allow novel
applicability with good solubility in organic solvents, low high-performance materials with synergistic performance
melting point and exceptional blending compatibility [9, [22].
10]. Binary PLA/PCL blends offer advancements in terms However, very little attention has been devoted to the
of physical properties and biodegradability. Amorphous research of multicomponent polymer blends from biopoly-
PLA, having high degradation rate, allows tensile strength mers. Compatible properties among PLA, PCL and TPS sug-
improvement, whereas crystalline-rubbery PCL allows gest biodegradable blends with good performance can be
toughness improvement but has much lower degradation obtained, such as appropriate processing and melt strength,
rate [11–13]. Broader packaging applicability is hindered strength and toughness balance and required thermal proper-
by their inherently high price, where starch was introduced ties. PLA and PCL were used as a matrix and TPS was added
as low-cost natural biopolymer. to accelerate biodegradation as well as to reduce the rela-
Native starch, one of the most abundant and widespread tive cost. Thereof, morphology, crystallinity, surface tension,
natural food sources, is a main component of many plants mechanical and thermal properties as well as water absorp-
such as corn, potato, rice, cassava, etc. Starch shows bio- tion tests of binary and ternary blends were investigated and
degradability, low density, nonabrasive nature, but cost- discussed. For that purpose, blends with different binary
efficiency is the main reason promoting starch as one of (PLA/PCL) and ternary composition (PLA/PCL/TPS) were
the most attractive environmentally friendly biopolymer prepared using Brabender kneader blending followed by a
[14–16]. Hence, it is often used to produce blends with hydraulic press process. The aim of this work is to show how
non-biodegradable polymers and, more recently, with bio- the development of multicomponent biodegradable polymer
degradable materials. Starch is not thermoplastic material. blends (ternary PLA/PCL/TPS blends) prepared by simple
Native starch generally exists in a granular state due to the melt blending can support viable packaging materials.
inherent hydrogen bond between adjacent molecules. The
transformation of starch into bioplastics for all applications
requires starch gelatinization i.e. the disruption of granules Materials and Methods
by applying mechanical shear stresses and heat in the pres-
ence of suitable plasticizers, usually glycerol and water, Materials
under specific conditions. Thereof, thermoplastic starch
(TPS) biopolymer is produced [17, 18]. TPS also shows The native wheat starch used was Srpanjka made up of 23%
some disadvantages that restrict its wider applications, such amylose and 77% amylopectin obtained from Agricultural
as dominant hydrophilic character, high degradation rate and Institute, Osijek, Croatia. Glycerol was supplied by Gram
poor mechanical properties, particularly under wet condi- Mol, Zagreb, Croatia. Polylactide (PLA) pellets were sup-
tions [19]. The best way to overcome these drawbacks is plied by Nature Works LLC, USA (Ingeo™ Biopolymer,
to blend TPS with other biodegradable polymers. As men- 4043D). PLA melting point is in the 145–160 °C range
tioned before, blending polymers is a simple, rapid and with a glass transition temperature (­ Tg) between 50–70 °C
cheap method suitable for achieving combinations of prop- with density of 1.24  g  cm −3 at 25  °C, and molecular
erties that are generally unattainable by a single polymeric weight around 110.000 g mol−1. Polycaprolactone (PCL)

13
Journal of Polymers and the Environment

pellets were supplied by Sigma-Aldrich, Germany (Poly- of biodegradable PLA/PCL/TPS ternary blends with their
caprolactone 440,744-500G). It exhibits an average Mn of chemical structures.
70.000–90.000 g mol−1 by GPC, Mw/Mn < 2, and density
1.145 g cm−3 at 25 °C. According to our research the melting Methods
point of PCL is located at 58.6 °C whereas the glass transi-
tion temperature of PCL is at − 62.8 °C [23]. The melt flow index, MFI, of the polymers, binary and
ternary blends was determined using an apparatus (Zwick
4100, Germany) with a capillary die of 8 mm length and
Preparation of Binary and Ternary Blends 2 mm diameter. Seven grams of each sample was loaded into
the barrel of the apparatus and then heated to the specified
The TPS was prepared by a melt extrusion process with temperature of 190 °C. A driving weight of 2.16 kg for the
glycerol as a plasticizer. The ratio of plasticizer to wheat material was applied to the plunger and the molten sample
starch was 30:100 (w/w). The TPS were extruded in a was forced through a die. The extrudates were collected and
laboratory scale Brabender single-screw extruder (Model weighed and MFI values were calculated in g (10 min)−1.
19/20DN; Germany; screw ratio L/D = 1:1 and die 4 mm) The morphology of the blends was observed from the
with a screw speed of 40 rpm and dosing speed of 15 rpm. cross-sections of cryogenically fractured surfaces of the
The temperature profile of the extruder barrel from the feed moulded bars by Vega EasyProbe 3 scanning electron
zone to the die was 100, 100 and 130 °C. The extruded microscope (Tescan, Brno, Czech Republic), SEM, with an
samples were air-dried overnight and then conditioned in a accelerating voltage of 15 kV. The samples were cooled in
sealed polyethylene bags at room temperature to maintain liquid nitrogen then fractured into specimens which were
stable moisture content. separately mounted on a stub using a conductive tape and
Briefly, the binary PCL/PLA and ternary PCL/PLA/TPS sputtered with gold prior to fractographic examination.
blends were blended for 6 min using a Brabender kneading The contact angle measurements were performed on a
chamber at 170 °C with a rotor speed of 50 rpm. Firstly, the contact angle goniometer, DataPhysics OCA 20 Instrument
binary PCL/PLA blends were blended and then TPS was GmbH. The measurements were carried out at a tempera-
blended with them in a constant weight composition of 30 ture of 23 °C ± 0.2 °C and a relative humidity of 60% with
wt. %. The mass ratios of PCL to PLA were s 100/0, 70/30, a volume drop of 2 μL with a stabilization period of 10 s,
50/50, 30/70 and 0/100. PCLx/PLAy was used as the symbol using an injection system. The drop shapes were geometri-
for the PCL/PLA blend, where x and y indicated mass per- cally determined from Laplace–Young fitting integrated
centages of PCL and PLA, respectively. The preparation of into the software (SCA20, version 2.01). The average val-
binary PCLx/PLAy blends was followed by the preparation ues of at least five drops of the same samples at difference
of ternary blends where binary blends were selected as the places were taken and the standard deviation was less than
polymeric matrix (PCLx/PLAy/TPS). 2% for all samples. The contact angle of each sample was
Subsequently, after homogenization, the blends were measured using three different liquids (water, formamide
compression moulded at 25 bar and 180 °C for 5 min using and diiodomethane). Distilled water, ­H2O, (twice distilled
a laboratory hydraulic press machine Fontune, Holland. λ = 1.33 µL cm−1) and formamide, ­CH3NO, (p.a. 99.5%,
Also, the neat PLA and PCL were subjected to the mixing Fluka) were used as polar solvents, while diiodomethane,
treatment in order to have the same thermal regime as the ­CH2I2, (p.a. 99%, Aldrich) was used as non-polar solvent,
blends. Scheme 1 displays an illustration of the preparation respectively. The literature values of their surface energies

Scheme 1  Illustration of the preparation of biodegradable ternary PLA/PCL/TPS blends

13
Journal of Polymers and the Environment

are [24]: ­H2O: γp = 51.0  mJ  m−2 and γd = 21.8  mJ  m−2; rate of thermoplastic materials. Interfacial characteristics as
­CH3NO: γp = 19.0 mJ m−2 and γd = 39.0 mJ m−2; ­CH2I2: a consequence of developed microstructural and structural
γp = 0.0 mJ m−2 and γd = 50.8 mJ m−2. parameters consequently influence the flow ability of poly-
The X-ray diffraction, XRD, was conducted using Shi- mer blends, so knowing MFI of a blend indirectly points out
madzu diffractometer XRD 6000 with CuKα radiation at to dispersion and interaction between the polymers. It should
acceleration voltage of 40 kV and current of 30 mA. Data be noted that MFI provides only a very crude approxima-
were collected between 2–70°2θ in a step scan mode with a tion. Still, obtaining MFI is heavily utilised in industry for a
step 0.02°2θ and counting time of 0.6 s. The degree of crys- quality control purpose before subjecting polymers to further
tallinity ­(Xc %) was estimated on a behalf of integral areas machine processing [25]. Here, the MFI value of the ther-
of crystalline and amorphous contribution (Pseudo-Voight moplastic starch could not be determined, pointing out to a
fit) upon baseline correction (polynomial fit). polymer that is not suitable for machine processing due to
Thermal gravimetric analysis, TGA, of the blends was blockage occurrence. Processing inability was stressed out
evaluated using thermogravimetric analyser TGA Q500 as one of the disadvantages of starches. MFI values were
(TA Instruments, New Castle, USA). Samples of approxi- obtained for PCL and PLA. From the Fig. 1 it is obvious that
mately 10 mg were heated from 25 to 600 °C at a constant the MFI values of PLA are much lower than those of PCL,
heating rate of 10 °C min–1 under a nitrogen atmosphere which may be due to higher molecular weight. MFI values
(60 ml min–1). The initial degradation temperature ­(T5%) was suggest that the incorporation of PCL in PLA significantly
calculated at 5% mass loss, while temperatures at the maxi- simplifies the processing of PLA/PCL blend due to molten
mum degradation rate (­ Tmax) for each stage were determined PCL properties.
from the first derivatives of the TGA curves (DTG). The MFI values of PLA/TPS and PCL/TPS binary blends
Water absorption, WA, of samples was determined indicate a uniform distribution of starch in PCL and PLA
by preparing 2 × 2 square inch thin film specimens using based matrices. The increase of MFI value indicates that
hydraulic press, Fontune, Holland, followed by hot press the viscosity of the blend is lower than thermoplastic starch
process for 4 min at 160 °C and 500 kPa and finally dried value. This is due to homogeneous blending of the wheat
overnight in a desiccator. The specimens were weighed every starch that has irregularly shaped particles, which increases
24 h for the next 8 days (192 h). Firstly, the specimens were the space between molecules and therefore diminishes the
removed from water then blotted dry with paper, weighted strength of intermolecular forces between polymer coils.
and returned to the distilled water again (23 °C, 100% RH). Possibility the blending process enhanced the formation
WA of each specimen was calculated as follows (1): of particle-matric interaction, causing the blend to become
m1 − m0 smoother and hence increase the mobility. It is evident that
WA = (1) MFI decreases with the content of thermoplastic starch in
m0
binary and ternary blends, which is explained with increas-
where m1 is the weight of the specimen at the specific time ing viscosity of the blends, higher molecular weight and
interval and m0 is the initial dry weight of the specimen. higher elasticity, all consequently diminishing the polymer
Tensile test of the blends was performed using a univer-
sal testing machine, Zwick 147670Z100/SN5A according
to standard ISO 527. Tests were performed at ambient tem-
perature (23 °C) and a cross-head speed of 50 mm min−1
to determine the tensile strength, elongation at break and
elastic modulus. In all mechanical tests, the reported data
are the mean and standard deviation from at least 5 measure-
ments. All samples were conditioned for 72 h at 23 °C with
a relative humidity of 50%.

Results and Discussion

Melt Flow Index

The melt flow index (MFI) is widely accepted parameter


that reposes on molecular weight and viscosity values of
the tested materials to describe the uniformity of the flow Fig. 1  Melt flow index for investigated samples

13
Journal of Polymers and the Environment

flow ability [25, 26]. On the other hand, the addition of PCL Table 2  Interfacial tension of Polymer pairs Interfacial
and PLA to thermoplastic starch increases the thermoplastic the polymer pairs tension
starch MFI value. Such behaviour is particularly interesting, (m N m−1)
as thermal processing of starch-based polymers is known to PLA/PCL 8.7
be demanding. However, with the decrease of PCL viscos- PLA/TPS 1.5
ity, PCL/TPS blend turns out to have higher MFI value than PCL/TPS 15.5
PLA/TPS blend.

Interfacial Properties
angle values with water (PLA = 68.0°, PCL = 99.7° and
The properties of polymer blend depend on composition and TPS = 55.6°) confirms that the PCL surface was the most
structure of individual components and also on interfacial hydrophobic with the highest contact angle value with water.
interactions between polymers [1, 7, 17]. Measurements of PLA shows a slightly lower contact angle value than PCL,
contact angle are widely used for the evaluation of material indicating a slightly less pronounced hydrophobic character
surface characteristics, such as wettability, hydrophilicity of PLA. On the contrary, low value of contact angle between
and surface energy. Thus, interfacial tension plays an impor- water and TPS surface is a consequence of polarity, i.e. pro-
tant role in the control of the hierarchal morphology of ter- nounced hydrophilic nature of TPS. The highest value of
nary blends. The miscibility of components into blends can polar component is observed for TPS; the lowest for PCL,
be predicted by surface free energy calculations using vari- whereas dispersive component is in very narrow interval
ous models and theories. Surface free energy characteristics from 22.8 to 24.7 m N m−1 for these polymers. The low
of examined materials were investigated via contact angle value of polar component as of PCL indicates nonpolar PCL
measurements with different test liquids having specified surface, i.e. possibility of establishing interaction at inter-
values of surface free energy. The contact angles of indi- phase mainly by dispersive forces.
vidual biodegradable materials (PLA, PCL, TPS) with water, Therefore, it is expected that the presence of TPS in
formamide and diiodomethane are listed in Table 2. Surface binary blends will change the surface wettability and
tensions of the samples were calculated using a model of hydrophilicity due to the strong intermolecular hydrogen
harmonic average from the measured contact angles (θ) and bonding at the film surface which causes surface polarity
the values of the interfacial tension were calculated by the of biopolymer films. Monitoring the development of mor-
equation of Wu (Eq. (2)) [27]: phology with respect to theoretical spreading coefficients
( d d p p
) requires knowing of the interfacial tensions for all binary
𝛾1 𝛾2 𝛾1 𝛾 2 blends; PLA/PCL, PLA/TPS and PCL/TPS (Table 2).
𝛾12 = 𝛾1 + 𝛾2 − 4 d + p (2)
𝛾1 + 𝛾2d 𝛾1 + 𝛾2
p
Understanding the mechanisms behind strong surface
adhesion development in binary blends can indicate which
where θ is the contact angle, γ is the surface tension, γd is the interface could exhibit stronger adhesion in ternary sys-
dispersion component and γp is the polar component. γ12 is tems. The highest value of interfacial tension is obtained
the interfacial tension between materials 1 and 2, γ1 and γ2 for PCL/TPS blends (γ PCL/TPS  = 15.5  mN  m −1 ) and it
are the surface tensions of the two contacting components indicates the lowest interaction extent between these two
in the blends. The values of calculated parameters are given polymers, i.e. poor miscibility of PCL and TPS polymers.
in Table 1. On the contrary, the lowest value of interfacial tension is
It is known that lower contact angle with water is char- for PLA/TPS polymers, which indicates good adhesion
acteristic for materials with higher surface wettability and between PLA and TPS polymers (γPCL/TPS = 1.5 mN m−1)
hydrophilicity. Respectively, PCL surface shows good wetta- and confirms better interactions are established between
bility and hydrophilicity. The comparison of average contact

Table 1  Contact angle and Sample Contact angle (°) Surface tension (m N m−1)
surface tension of the polymers
Water Formamide Diiodomethane Total (γ) Dispersion com- Polar
ponent (γd) compo-
nent (γp)

PLA 68.0 72.5 65.2 38.1 22.8 15.4


PCL 99.7 83.1 70.8 27.0 24.2 2.8
TPS 55.6 55.4 63.8 47.4 24.7 22.7

13
Journal of Polymers and the Environment

those phases. The value of interfacial tension for PCL/PLA


pairs is between afore mentioned values (8.7 m N m−1).
Other approach to theoretical prediction of morphology
development in ternary blends reposes on Harkin’s spread-
ing coefficients [28]. The advantage of this approach is that
the knowledge of the interfacial properties is sufficient for
prediction of the morphology development. Based on the
calculated interfacial tension data, it is possible to predict
the morphology of multiphase polymers [28, 29]. Using
the values of the interfacial tensions given in Table 2, it is
Fig. 2  Possible phase morphologies for a PLA/PCL/TPS polymer
possible to calculate the set of three spreading coefficient blend, and their relations of the spreading coefficients [28, 30]
for a PLA/PCL/TPS blends as follows:
( )
λPLA/PCL/TPS = γPLA/TPS − γPLA/PCL + γPLA/TPS = −26.8 Structural Characterization of the Blends, XRD
(3) Analysis
λPCL/PLA/TPS = γPCL/TPS − γPLA/PCL + γPLA/TPS = 5.5 (4)
( )
All blends were analysed by means of X-ray diffraction,
( ) where the constituent phases and blends were shown sepa-
λPLA/TPS/PCL = γPLA/PCL − γPLA/TPS + γPLA/TPS = −9.2 rately (Figs. 3 and 4). The constituents were identified;
(5) yet the ICDD diffraction assignation is not meaningful for
where λPLA/PCL/TPS are the interface tensions between poly- compounds having low structural order. For PCL, several
mers. γPLA/TPS is the interfacial tension between PLA and characteristic peaks could be easily observed, as this com-
TPS phases, γPCL/TPS is the interfacial tension between pound displays semi-crystalline behaviour. Afterwards,
PCL and TPS phases and γPLA/PCL is the interfacial ten- this will be used to discuss the degree of crystallinity.
sion between PLA and PCL phases. A positive λPLA/PCL/TPS The crystallite sizes of PCL remain similar for all blending
indicates the tendency of component PCL to spontaneously conditions. On the contrary, PLA is predominately amor-
spread at the interface of components PLA and TPS, which phous, whereas only single characteristic low intensity
is considered as the complete wetting regime. However, broad peak may be identified. For native wheat starch,
when all the three coefficients are negative, partial wetting semi-crystalline behaviour is visible as several characteris-
is predicted at triple-phase contact where all phases meet, tic peaks arise. Starch thermoplastic derivative, biodegrad-
since the spreading does not occur. A negative but close to able polymeric TPS material resembles semi-crystalline
zero spreading coefficient suggests a weak partial wetting to almost amorphous behaviour, with possible occurrence
driving force, whereas the negative spreading coefficient of one or more characteristic peaks. PCL shows the char-
implies a strong driving force towards partial wetting. Two acteristic crystalline pattern (Fig. 3) with peaks at about
wetting regimes are defined based on the spreading coef- 21.5, 21.8, 23.7 and 36.1°2θ attributed to the (110), (111),
ficient; complete wetting and partial wetting. For positive (200) and (213) reflections, respectively [31–33]. PLA is
spreading coefficients, the liquid drop completely spreads characterized by one weak peak centred at about 17°2θ
and forms a thin-film on the substrate to minimize the sur- within broad amorphous hump centred at about 20°2θ,
face energy. However, the liquid drop does not spread, rather which consequently does not allow monitoring of possible
forms a spherical cap with an equilibrium contact angle on structural transitions [34, 35]. TPS shows characteristic
the substrate when the spreading coefficient is negative [29]. peak at about 20°2θ (Fig. 3), attributed to the processing-
In our case, the results of interfacial tension indicate induced crystallinity of polyols molecule. Polyols, i.e.
that in PLA/PCL/TPS ternary blends, the interfacial ten- glycerol also shows considerable influence at the TPS
sion of PCL/TPS is much greater than that of PLA/PCL, structural parameters. Commonly, starch is plasticized by
whereas the interfacial tension of PLA/TPS is lower than adding an appropriate amount of plasticizer, where a mate-
tension of PLA/PCL. A multicomponent polymer system rial flexibility is increased by reduction of intramolecu-
will usually develop a morphology having lowest free lar hydrogen bonding, and thereof allows intermolecular
energy; therefore in this system TPS phase is most likely spacing along polymer chains. Thereby we can expect an
encapsulated by PLA phase in PCL matrix. Figure 2 illus- increase of molecular mobility and a decrease of crystal-
trates possible phase morphologies for PLA/PCL/TPS linity [12, 36, 37]. These single-phase characteristics were
polymer blends and their relations of the spreading coef- used to identify and discuss the behaviour of the blends.
ficients according Fu et al. [30]. The behaviour of PCL in blends with PLA is consistent
with the content of the sole constituents (Fig. 3a). Blending

13
Journal of Polymers and the Environment

Fig. 3  a XRD diffractograms


of the binary PCL/PLA and
b XRD diffractograms of the
ternary PCL/PLA/TPS blends

Fig. 4  a Overlap of the XRD


diffraction patterns of the PCL/
PLA binary blends, where com-
positional changes reflecting
in quantitative changes may be
observed and b overlap of the
XRD diffraction patterns of the
PCL/PLA/TPS ternary blends

biodegradable and ductile PCL with PLA should allow novel the miscibility, among other techniques, the addition/incor-
materials showing upgrades in already good mechanical poration of a third component such as small molecule should
properties of PLA yet with retained biodegradability on reduce the size of the dispersed phase and thereby improve
behalf of PCL [38, 39]. The blending may be hindered by toughness and elongation at break of the overall blend. From
thermodynamic immiscibility of PLA and PCL. To enhance the available literature, the mechanisms of the ternary blends

13
Journal of Polymers and the Environment

formation are not completely clear, especially in the case features in general follow the relative compositional ratios
of TPS addition. The PCL/PLA/TPS blend is prepared to of the constituents.
allow better mechanical properties, especially rubbery elas- Again, critical information obtained from the XRD data
tic properties [40]. The addition of TPS should not decrease is the observation of influence on the structure through
the abovementioned performance. Given their compatible the degree of crystallinity. The addition of TPS into PLA/
properties, we believe that the polymers could be blended PCL blends suggests miscibility improvement was limited.
more successfully. Here we address these aspects only to Namely, the decrease of PCL relative crystallinity values
observe minimal, but unneglectable changes on behalf of (hindered crystallization) in the PCL-based blends points
diffraction data. Basically, the increase in the content of the out to phase separation process in ternary blends (Table 3).
predominately amorphous PLA within the semi-crystalline
PCL decreases overall peak, i.e. signal intensity proportion- Blend Morphology
ally to the ratio of the constituents. The distribution of the
polymeric domains, i.e. the level of the structural organiza- The fracture surfaces of pure thermoplastic starch, polycap-
tion of the semi-crystalline PCL blends is barely affected. rolactone and polylactide, as well as prepared binary and
The PCL/PLA in blends with TPS follows the behaviour of ternary polymer blends, examined by a scanning electron
the PCL/PLA blends (Fig. 3b). Namely, from the structural microscopy (SEM) are shown in Figs. 5–7. Before SEM
point of view, the addition of TPS only slightly interferes analysis, surfaces were sputtered with gold to ensure the
with the mechanisms of the ternary blend formation as well conductivity of the samples. Figure 5a shows a SEM micro-
as with the structural level of ordering. First, the change in graphs of native wheat starch before plasticization where
overall XRD signal is a proportional function of the relative granular microstructure is visible. Granules are more or
constituents’ contribution. Overlap of the XRD diffracto- less isotropic with broad size distribution and significant
grams of PCL/PLA binary blends and PCL/PLA/TPS ter- agglomeration. The texture of the granules is pore and crack
nary blends allowed observation of critical differences in free, surface roughness is low, edges are smooth and blunt.
the blend formation between the two blends (Fig. 4a). While During the plasticization process, granular morphology
the correct calculation of the crystallite size from diffraction went through a transformation yielding homogeneous poly-
signal is not possible for polymeric materials, the correlation meric film. The degradation of hydrogen bonding between
of the PCL diffraction peaks at least reveals the size of the the starch molecules occurred synchronously with the for-
PCL crystallites remains barely unchanged for samples with mation of the hydrogen bonds between the plasticizer and
and without TPS. starch molecules, whereas thermoplastic starch was obtained
Diffraction peaks observed for PCL at 21.3 and 23.6°2θ, (Fig. 5b). Therefore, SEM microphotograph of TPS (Fig. 5b)
attributed to the (111) and (211) crystal planes indices [33], shows uniform texture and morphology, with no unreacted
remain in the blended specimens. Regarding peak shifting, starch residuals. The absence of wheat starch granular resid-
in the PCL/PLA/TPS blends, all major PCL peaks; (110), uals suggests that the plasticization process was complete.
(111) and (200) plane reflections, remain constant 2θ posi- It can be concluded that the conditions for obtaining TPS
tions, suggesting stable lattice of PCL. Some authors sug- were well selected to transform granular wheat starch into
gest PCL crystal lattice extension or contraction may indi- homogeneous TPS over plasticization process (Fig. 5b).
cate good PCL-TPS interaction, which is not the case here SEM micrograph of fracture surface of the pure PLA
[31]. Another way of structural changes, such as hindered (Fig. 6a) shows flat and smooth surface, corresponding to
crystallization on behalf phase separation, points out to lim- low energy dissipation from crack initiation to propagation.
ited homogeneity between TPS and PCL and insignificant This corresponds to typical fracture surface of brittle mate-
improvement of PCL-TPS interaction. Also, the differences rial with low roughness and absence of plastic deformation,
on behalf of TPS addition are limited only to minor peaks
that may occur also due to starch derivatives (TPS) degra-
Table 3  Crystallinity of the PCL in the binary PCL/PLA and in the
dation. Namely, TPS is prone to water adsorption. Obvi- ternary PCL/PLA/TPS blends
ously, PCL/PLA will reduce the overall blend susceptibility
to water adsorption, as the constituents are known for good Composition Crystallinity Composition Crystallinity
permeability with slow hydrolysis [41, 42]. The addition PCL 33.7 PCL/TPS 22.7
of TPS will show only trace changes of water based deg- PCL70/PLA30/ 22.2 PCL70/PLA30/ 17.1
radation. Finally, in Fig. 4b the XRD diffractograms of the TPS TPS
PCL/PLA binary blends and PCL/PLA/TPS ternary blends PCL50/PLA50 15.7 PCL50/PLA50/ 13.7
are shown where compositional changes reflecting in quan- TPS
titative changes may be better observed. The quantitative PCL30/PLA70 11.9 PCL30/PLA70/ 8.8
TPS

13
Journal of Polymers and the Environment

Fig. 5  SEM micrographs of: a


native wheat starch, b thermo-
plastic starch

Fig. 6  SEM micrographs of: a


PLA, b PCL, c PLA/TPS and d
PCL/TPS

which is correlated with the properties of PLA [5, 7, 11]. elongated fibril features, that can be observed from SEM
Meanwhile, SEM micrograph of the PCL (Fig. 6b), where micrograph (Fig.  6b) points out to ductile deformation
certain degree of structural order is observed, shows no behaviour of PCL.
traces of no brittle fracture, indicating that a certain degree The addition of thermoplastic starch to PLA and PCL
of material elasticity is present. This is also obvious from clearly shows the completely different morphology, as
high elongation at breaking point. In addition, presence of shown in Fig. 6c, d. Comparing the SEM micrographs of

13
Journal of Polymers and the Environment

Fig. 7  SEM micrographs of binary a PCL70/PLA30, b PCL50/PLA50, c PCL30/PLA70, and ternary, d PCL70/PLA30/TPS, e PCL50/PLA50/
TPS, and f) PCL30/PLA70/TPS blends

binary PLA/TPS and PCL/TPS blends, a better miscibility mechanical properties of such blends. Due to unsatisfying
is observed in PLA/TPS blend (Fig. 6c). SEM micrograph of transfer of stress from the matrix to the dispersed particles
the PCL/TPS (Fig. 6d) blend shows a higher accumulation of and vice versa, cracks at the interface could occur. Further-
TPS on the fracture surface, which is easier to squeeze out more, with the increase of PCL content in PCL/PLA blends,
when breaking, which indicates poor penetration of thermo- the fibril microstructure of PCL is more pronounced, while
plastic starch into the PCL matrix and the absence of interac- in contrast, with higher PLA content SEM micrographs
tions that would contribute to better miscibility of PCL and show isotropic morphology.
TPS. The formation of agglomeration-prone TPS flakes can SEM micrographs of ternary PCL/PLA/TPS blends
be clearly observed, which were mechanically induced to (Fig. 7b, d, f) show that TPS addition results in poor adhe-
form blocks and accumulate. Since PCL is a highly hydro- sion at the phase interface and the presence of micropores
phobic biopolymer and TPS is highly hydrophilic, the result and cracks in the polymer matrix. By adding TPS to the
is a strong phase separation resembling to two-phase mor- PCL/PLA blends (PCL/PLA/TPS), miscibility or compat-
phology of the PCL/TPS blend. ibility of TPS was not improved. In contrast, the accumula-
Binary PCL/PLA blends (Fig. 7a, c, e) exhibit rough sur- tions of extruded starch and consequently residual cavities
face and bimodal morphology, characterized by the exist- are observed after extrusion of starch on the fracture surface.
ence of various voids and growth of fibril structures with The best morphology was achieved with a PLA content of
the increase of PCL content. This indicated poor adhesion 70 wt. %, or with purely PLA matrix (Fig. 7f). Majority of
between PCL and PLA phases and high interfacial tension, starch granules are well covered by the PLA matrix, and
which can result in poor functional properties of such PCL/ surface fracture probably occurred because of PLA matrix.
PLA blends. There is also a clear, sharp interface between It is also possible to notice some degree of material elastic-
PCL and PLA, which indicates poor penetration of either ity, on behalf of PCL presence in the PCL70/PLA30/TPS
PCL or PLA, which is a consequence of the poor adhesion blend (Fig. 7b).
at the interface. This leads to limitations in stress trans-
fer through the system which could result in diminished

13
Journal of Polymers and the Environment

Thermogravimetric Analysis for PLA. PCL is the most stable polymer with the onset of
thermal decomposition at a temperature of 364.3 °C with a
In order to evaluate thermal stability, thermogravimetric maximum degradation rate at 408.1 °C and completion of
(TG) analyses of constituents PCL, PLA, TPS, their binary degradation at 441.8 °C. Degradation of PCL occurs in a sin-
and ternary blends were performed in inert (nitrogen) gle stage thermal interval about 22 °C wider in comparison
atmosphere. Initial decomposition temperature, T ­ 5%, final with PLA and TPS (Fig. 8b).
decomposition temperature, ­Tf, maximum decomposition According to the Persenaire et al. the thermal degrada-
temperature, ­Tmax, with associated mass loss ratio, Δm, tion mechanisms of PCL is characterised by two stages that
decomposition temperature interval, ΔT, and char residue occur simultaneously [44]. In the first stage of degradation,
at 600 °C, R
­ 600 °C, were determined from TG and derivative cis-elimination occurs, i.e., the bursting of polyester chains
TG (DTG) curves (Table 4). Initial decomposition tempera- via the reaction of pyrolysis of esters, with the release of
ture depends on blend composition and indicates in thermal ­CO2, ­H2O and the formation of acidic carboxyl groups. In
stability of the blends. Figure 8 shows the TG and DTG the second stage of degradation, specific breakage of the
curves of constituent polymers as well as binary PCL/TPS chain-end occurs, resulting in caprolactone (cyclic mono-
and PLA/TPS blends, respectively. Whereas thermal stabil- mer) due to the depolymerization process. However, the
ity of PCL is not high due to low melting and glass transition initial blocking of the chain-end hydroxyl groups may not
temperatures, PCL shows highest thermal stability among completely prevent the occurrence of specific bursting of the
biodegradable polymers used in this study [8, 10, 23, 43]. chain end. It is therefore suggested the two processes occur
Figure 8 shows that thermally induced changes for PCL simultaneously. On the other hand, from the obtained PLA
start at 364.3 °C, which is higher than 322.6 °C observed curves (Fig. 8b) it can be seen that the thermal degradation

Table 4  Results of thermal properties


Sample T5% (°C) T1max (°C) ∆m1 (%) T2max (°C) ∆m2 (%) T3max (°C) ∆m3 (%) Tf (°C) ∆T (°C) R600°C (%)

PCL 364.3 – – – – 408.1 99.0 441.8 77.5 0.00


PLA 322.6 – – 367.6 98.3 – – 377.6 55.0 0.87
TPS 279.3 318.4 65.9 – – – – 338.1 58.8 10.08
PCL/TPS 291.6 314.3 18.9 – – 406.1 70.9 429.8 138.2 2.41
PLA/TPS 221.2 243.4 17.0 327.4 70.0 – – 341.7 120.5 5.56
PCL30/PLA70 325.9 – – 365.2 75.4 398.2 23.2 418.3 92.4 0.19
PCL30/PLA70/TPS 296.5 318.6 27.5 329.7 34.1 372.1 24.9 402.4 105.9 4.37
PCL50/PLA50 318.8 – – 355.1 44.3 408.0 53.3 422.0 103.2 1.09
PCL50/PLA50/TPS 294.5 311.2 25.0 356.3 29.5 401.8 36.4 424.1 129.6 2.90
PCL70/PLA30 333.4 – – 360.5 15.1 455.7 82.7 477.1 143.7 2.30
PCL70/PLA30/TPS 291.2 310.9 18.3 354.2 20.4 401.8 50.7 420.7 129.5 3.05

Fig. 8  TGA (a) and DTG (b) curves of PCL, PLA, TPS and binary PCL/TPS and PLA/TPS blends

13
Journal of Polymers and the Environment

of PLA also occurs in a single stage, in a relatively narrow rate temperature of PCL ­(T1max) and TPS ­(T2max) remain
temperature interval of about 55 °C. The onset of PLA deg- unchanged, indicating that PCL and TPS are immiscible
radation is at 322.6 °C, with a maximum degradation rate (Table 4). The addition of thermoplastic starch to polylac-
at 367.6 °C, with a mass loss of 93.3% attributable to the tide significantly decreases the temperatures of maximum
cleavage of the PLA end macromolecules. Literature sug- degradation rate, which indicates a certain degree of interac-
gests that thermal decomposition of PLA occurs via com- tions are present, i.e. a certain degree of miscibility, which
plex mechanism which prevents simple obtaining of activa- corroborates well with SEM microphotographs (Fig. 2b).
tion parameters [45, 46]. Namely, there are many factors Accordingly, the thermal behaviour of the PLA/PCL blend
that affect the thermal stability of polylactide, among other was found to be intermediate to those of constituent poly-
moisture, hydrolysed monomers and oligomers, molecular mers (Table 4). The thermal decomposition of the PLA/PCL
weight and residual metals [47]. The thermal degradation of blends occurs in two stages. The first stage is between 300
PLA begins with the hydrolysis of water that has been left and 370 °C, dominated by thermal decomposition of the
behind during processing, followed by a depolymerization, PLA since its degradation onset is found at a lower tem-
random degradation of the main chain, followed by intra- perature, while the second stage of decomposition in the
molecular and intermolecular transesterification [48]. Then, range of 370–440 °C relates to the thermal decomposition
the starch chains begins to degrade at about 230 °C mainly of PCL as a more thermally stable component in the blend
due to dehydration of hydroxyl groups and the subsequent (Fig. 10). The TGA and DTG curves clearly show all the
formation of unsaturated and aliphatic low molecular weight stages of thermal degradation for PLA and PCL are shifted
carbon species [15]. to higher temperatures, due to the increase of PCL content
Unlike single degradation stage of PCL and PLA, the in the blend. The exception is observed for a blend with 50
thermoplastic starch degradation occurs in two stages. The wt. % of polycaprolactone added, where the decomposition
TGA curve of TPS (Fig. 8a) shows mass loss up to 250 °C onset temperature is slightly lower compared to the pure
associated with the evaporation of water, glycerol and other PLA. For ternary blends, the addition of TPS to the PCL/
volatile compounds which are present in TPS due to its PLA blends decreases the onset temperature of degradation
extreme hydrophilic character. Evaporation is followed by up to about 30 °C (due to the lower thermal stability of the
thermal decomposition of the polysaccharide, i.e. starch TPS). The extent of mass loss in the temperature range from
between 250 and 350 °C with a maximum decomposition 30 to 300 °C relative to the composition of the blends was
rate at temperature of 318.4 °C. The thermal decomposition also compared. A similar influence of TPS on the thermal
of starch is primarily related to the breakage of strong bonds, stability of blends has been obtained by other researchers [4,
such as glycosidic bonds. The process of thermal dehydra- 17, 19, 49]. The initial mass loss up to 70 °C is attributed
tion follows, with gaseous CO, ­CO2, ­H2O, aldehydes, alco- to the evaporation of water, and other volatile compounds
hols, acids, aliphatic and furan fragments as decomposition present in TPS due to its pronounced hydrophilic character
products [15]. The possible starch degradation mechanism is barely noticeable in PCL/PLA/TPS blends, indicating that
according to Worzakowska is shown in Fig. 9 [16]. TPS small amount of water is adsorbed within their structure.
shows the onset of degradation at lower temperature than The TGA and DTG thermal decomposition curves of PCL/
PCL and PLA polymers, indicating its low thermal stability. PLA/TPS blends show three stages of thermal decomposi-
From the TGA/DTG results it can be seen that the addi- tion (Fig. 10). The first stage is related to the degradation of
tion of TPS lowers the degradation onset temperature, reduc- TPS as the least stable component in blends and the second
ing the thermal stability of each polymer. The addition of stage is related to thermal degradation of PLA. The last stage
TPS to PCL resulted in a decrease of ­T5% from 364.3 °C to of thermal degradation is attributed to the degradation of
291.6 °C while the thermal stability of PLA is reduced by PCL in PCL/PLA/TPS blends because of its highest thermal
100 °C (Table 4). Furthermore, the maximum degradation stability. The results confirm the TPS phase addition reduces

Fig. 9  Thermal decomposition of TPS [16]

13
Journal of Polymers and the Environment

Fig. 10  TGA (a) and DTG (b) curves of binary PCL/PLA and ternary PCL/PLA/TPS blends

the thermal stability of the blends. The weight percentage Water Absorption
of char residues for all samples is presented in Table 4. It
can be observed that TPS contributed to generating a higher The water absorption (WA) behaviour of the constituent
amount of residue due to the char residues of the neat TPS polymers, their binary and ternary blends as a function of
(10.08%). It may be because the starch is a carbohydrate, water immersion time is presented in Fig. 11. As aforemen-
which yields high charred products. Whereas the neat PCL tioned, one of the major disadvantages of TPS is its ten-
presented completely degradation without char residues, the dency for water absorption. The hydroxyl groups in starch
neat PLA piece resulted in very low char content of approxi- form a hydrogen bond with water, so any improvement in
mately 0.87%. By adding the TPS, this value increased up to water sensitivity is significant for many practical applica-
2.41% and 5.56% in the binary PCL/TPS, PLA/TPS blends, tions of starch-based materials [50]. On the other hand;
respectively. The binary PCL/PLA blends presented inter- polycaprolactone and polylactide have a pronounced hydro-
mediate char residue. This result can be related to additives phobic character and high resistance to water. Comparing
incorporated into the biopolymers by the manufacturer. As the results of water absorption behaviour for PCL and PLA
expected, ash content increased in ternary blends due to the (at a plateau point 0.8%), more pronounced hydrophobic
TPS presence. character is shown by polycaprolactone (at a plateau point
0.2%) (Fig. 11a).
Therefore, water absorption was investigated on behalf
of a difference in water affinity of TPS (hydrophilic) to PCL

Fig. 11  Water absorption versus immersion time for a PCL, PLA, PLA/TPS, PCL/TPS and b binary PCL/PLA and ternary PCL/PLA/TPS
blends

13
Journal of Polymers and the Environment

and PLA (hydrophobic). It is obvious that water absorption remain constant for about 4 days. Maximum water absorp-
of blends will increase if TPS is added. Namely, that water tion was 6.3% for ternary blends with higher content of PCL
absorption of the polymer blends depends on the weight (PCL70/PLA30/TPS) (Fig. 11). Water absorption of this
fraction of TPS, weight fraction of hydrophobic polymer, blend is typical for Fickian type of diffusion behaviour. In
the thermodynamic interaction energy between TPS and the Fick’s law, the concentration gradient is the driving force for
hydrophobic polymer, and their respective diffusion coeffi- diffusion and amount of the component diffused in a func-
cients [51]. The measuring of the water absorption for con- tion of time. Generally, water absorption process follows
stituent TPS was not possible in eight days because during the prediction of Fick’s law as the mass of water absorbed
the first day the TPS was saturated and broke down, which increases linearly with square root of time until equilibrium
confirmed its strong hydrophilic character. After 48 h, TPS plateau is reached.
turned into a paste that could not be removed from water The hydrophobic character of PCL and PLA causes
in order to measure weight. Placing the TPS in a high rela- reduction of water adsorption for ternary blends and thereof
tive humidity environment further reduced the strength of reduce their equilibrium water content. Likewise, when the
the samples to the point where they were no longer able to ratio of PCL or PLA in the blends increased, their water sol-
remain coherent. Therefore, the WA value for TPS was cal- ubility decreased, as shown in Fig. 11. The results confirm
culated based on the results during the first day and equals to the expected changes in the water affinity of the blends when
104.9%. In addition, the plasticizer itself contributes to this the ratio of PCL or PLA increased, which is an advantage of
extremely high WA value of thermoplastic starch. A plasti- these ternary materials.
cizer is necessary for processing starch into a thermoplas- For water absorption, it can also be noticed that an
tic material. Adding plasticizer has an impact on essential increase in PLA content reduces the plateau equilibrium
material properties such as water resistance, glass transition value; despite its slightly less hydrophobic character com-
temperature, and mechanical properties. Although glycerol pared to PCL. This behaviour can be explained by the better
is proven as the most common plasticizer for plasticizing miscibility of TPS and PLA and the absence of cavities as
starch, due to the high amount of end hydroxyl groups, it potential water retention sites.
pronounces the water sensitivity even more. We confirmed Observing the results of WA for binary PCL/PLA blends,
the thermoplastic starch is highly hydrophilic. regardless of their mutual composition, changes in water
As expected, the TPS sample shows the highest water adsorption are not specific. This result is expected due to
absorption due to its hygroscopic nature [8, 14, 17], whereas, both components having highly hydrophobic character.
the hydrophobic PCL and PLA samples showed the low- The absorption of excessive water weakens the TPS
est water absorption (Fig. 11). The water absorption in the properties. Thereof it is critically important to diminish the
binary TPS samples, increased linearly as a function of time property of starch plastics to absorb water. The pronounced
in the first 100 h (4 days). Finally, a plateau was observed, absorption of starch reposes on ability of water molecule to
indicating that water absorption had reached equilibrium easily diffuse through the starch network which also strongly
plateau. As reported previously for similar systems, the interacts with water molecules. Adding hydrophobic materi-
absorption is fast for shorter periods (t < 100 h) and slows als significantly reduces water absorption in concordance
down for extended reaction periods reaching an equilibrium with the theory for a polymer blend that predicts the dif-
plateau. The equilibrium swelling, or maximum relative fusion coefficient and water absorption of a hydrophilic
water uptake, or water uptake at equilibrium, corresponds polymer will be reduced when blended with a hydrophobic
to the plateau value observed at 100 h (4 days). polymer.
Furthermore, the incorporation of PCL and PLA can dra-
matically decrease the water absorption of TPS and reduce Mechanical Properties
the water sensitivity of the PCL/TPS (4.0%) and PLA/
TPS (2.5%) blends due to their hydrophobic characteristic Maintaining the quality of mechanical properties is greatly
(Fig. 11). Figure 11 shows that water absorption of PLA/ important because the material must be able to display stable
TPS blends are slightly lower than those of the correspond- performance under various loads. From the law of blends
ing PCL/TPS blend. This can be explained by a finer disper- for miscible polymers, the expected properties of the blend
sion of TPS phase in the PLA matrix due to its lower melt are near to the sum of properties of the constituents. For
viscosity, which leaves less free hydroxyl groups for water immiscible polymers, mechanical properties are markedly
binding. Additionally, the dense and homogenous micro- impaired. Results of mechanical properties: tensile strength,
structure can hinder the diffusion of water molecules in the elongation at break and elastic modulus of pure polymers,
PLA/TPS blends. binary and ternary PLA/PCL/TPS blends were evaluated
Ternary blends with thermoplastic starch show a sharp from stress–strain curves and the results are shown graphi-
increase of water absorption at the beginning and then cally in Fig. 12.

13
Journal of Polymers and the Environment

From the results of the mechanical properties shown in thermoplastic starch is about 5 times higher compared to
Fig. 12, the PLA is highly fragile with a very low elonga- pure PLA. The disadvantages of TPS can be mitigated by
tion at break values of 5.3%. Very low elongation at break, various modifications and by blending with polymers that
medium-to-high tensile strength of 45.3 N mm−2 and the have strong hydrophobic character. The blending produces
elastic modulus of ~ 1020.0 N mm−2 are responsible for PLA materials with low water sensitivity, high strength, and envi-
high brittleness. High strength, rigidity and low elongation able thermal stability with pronounced biodegradable prop-
at break of polylactide, nominate PLA as biopolymer appro- erties. Based on the obtained mechanical properties of each
priate for loading applications, such as fixing orthopaedic polymer (Fig. 12), it is expected that blending TPS with
accessories [52, 53]. On the other hand, PCL exhibits very biodegradable polymer, either PCL and PLA will improve
ductile behaviour, extensive elongation at break, capable to the tensile strength and elongation at break of TPS, whereas
undergo large deformations. It shows relatively low elastic reduce the elastic modulus, which will result in an improve-
modulus of 261.4 N mm−2, which limits its application for ment of the poor mechanical properties of thermoplastic
materials with high rigidity requirements (Fig. 12). Ther- starch. The mechanical properties of the pure polymers
moplastic starch exhibits extremely poor mechanical prop- obtained in this study corroborate well with the literature
erties compared to PCL and PLA, very low tensile strength values [5, 11, 20, 46, 54]. Some observed deviations were
of 4.1 N mm−2, low elastic modulus of 91.6 N mm−2, and not pinpointed, namely mechanical properties depend on
elongation at break 22.3%. TPS can be described as a highly morphology, molecular weight of the polymer, degree of
fragile polymeric material, i.e. material with strong hydro- crystallinity, plasticizer content in thermoplastic starch, its
philic character and poor mechanical properties, which composition as well as the source of the starch.
hinders its widespread use. Still, the elongation at break of

Fig. 12  Mechanical properties of binary PCL/PLA and ternary PCL/PLA/TPS blends a tensile strength, b elongation at break and c elastic
modulus

13
Journal of Polymers and the Environment

The addition of TPS to pure PCL and PLA polymers plasticization. Addition of TPS to PCL and PLA heavily
results in a decrease of the elongation at break of the affected their morphology, in favourable manner for binary
PCL while the elongation at break of the PLA increases PLA/TPS blend. Addition of TPS to PCL/PLA blends (PCL/
(Fig. 12b). Due to the extreme fragility of PLA, the addition PLA/TPS) resulted in poor adhesion at the phase boundary
of thermoplastic starch, due to the high content of plasti- and the presence of micropores and cracks in the polymer
cizer (glycerol), contributes to the increase in the elonga- matrix, where the immiscibility of the blends was clearly
tion at break. The tensile strength is the highest for PLA observed by the presence of two separated phases. The best
and the lowest for TPS, while the effect of blending TPS morphology was achieved with a PLA content of 70 wt. %,
in PLA is manifested by the decrease of tensile strength i.e. where the polylactide is a matrix (PCL30/PLA70/TPS).
(Fig. 12a). Also, the elastic modulus is the highest for PLA, Polymeric constituents and blends were successfully
and the lowest for TPS, while the addition of TPS results monitored using X-ray diffraction. From the structural
in a decrease of the elastic modulus value for PLA while point of view, the addition of TPS slightly interferes with
slight increase for PCL (Fig. 12c). Mechanical properties of the mechanisms of the ternary blend formation as well as
samples with TPS corroborate well with the SEM micropho- with the structural level of ordering. XRD allowed obser-
tographs, indicating their immiscibility. Researchers have vation of influence on the structure through the degree of
also shown that low adhesion between TPS and PCL occurs crystallinity. Namely, the decrease of PCL relative crystal-
due to poor interfacial interactions, which lead to a decrease linity values (hindered crystallization) in the PCL-based
in mechanical properties [49, 55, 56]. The addition of PCL blends points out to phase separation process in ternary
to PLA is manifested by a decrease of tensile strength and blends. Results of TGA analysis indicate that the addition
elastic modulus, while the elongation at break is increased of TPS decreases thermal stability of PCL/TPS and PLA/
compared to pure PLA, indicating a positive influence of TPS blends. In binary PCL/PLA blends, PCL contributes to
the addition of PCL on elongation at break of PLA/PCL the reduction of the elongation at break and elastic modulus,
due to its prominent flexibility (Fig. 12) [57, 58]. On the while increasing the tensile strength of PLA, on the other
other hand, the addition of PLA to PCL contributes to almost hand, the PLA contributes to tropeling of the stiffness of the
tripling of PCL/PLA stiffness. Figure 12 shows that the PCL. The addition of TPS into PLA/PCL blends impairs the
higher the proportion of PCL in PCL/PLA blends increases mechanical properties of the individual components in the
the elongation at break. Due to the extreme elasticity of the binary blends, suggesting miscibility was limited.
PLA, even a small increase of PLA-based blend stretching Finally, ternary biodegradable blends show great poten-
is extremely important for its application. It can be seen that tial as environmentally friendly materials primarily due to
the elongation at break of the PCL/PLA blend increases to significant cost reduction. Resolving waste accumulation is
392.6%, which is a remarkable contribution, compared to an issue that may be resolved by application of these ternary
pure PLA (5.3%). The elastic modulus of PCL/PLA blends biodegradable blends for disposable material. The topic may
reduces gradually by the increase of PCL content (Fig. 12c). need further research, focusing on eluting adequate com-
PCL/PLA blends show reduced brittleness due to the high patibilizer in order to achieve better adhesion of the blend
elongation at break value (Fig. 12b) [59–61]. constituents as well as improving miscibility and therefore
The addition of TPS in PCL/PLA blends reduces tensile obtaining better material properties.
strength, elongation at break and elastic modulus, impairing
mechanical properties. This occurs as physical incorporation Acknowledgements  This work was supported by projects PZS-2019-
02-1555 in Research Cooperability Program of the Croatian Science
of the TPS into the PCL/PLA matrix of the blend weakens Foundation funded by the European Union from the European Social
the London forces between the bland constituents. Another Fund under the Operational Programme Efficient Human Resources
reason for such behaviour is found in low molecular mass 2014-2020.
of TPS molecules, which can consequently withstand less
elongation at break compared to PLA and PCL. Author contributions  Conceptualization: [Vesna Ocelić Bulatović
and Dajana Kučić Grgić]; Methodology: [Vesna Ocelić Bulatović];
Formal analysis and investigation: [Vesna Ocelić Bulatović and Vilko
Mandić]; Writing—original draft preparation: [Vesna Ocelić Bulatović,
Conclusions Vilko Mandić and Antonio Ivančić]; Writing—review and editing:
[Vesna Ocelić Bulatović and Vilko Mandić]; Resources: [Vesna
Ocelić Bulatović and Dajana Kučić Grgić]; Supervision: [Vesna Ocelić
This investigation focused on preparing multiphase bio- Bulatović and Vilko Mandić].
degradable polymer materials by blending three promis-
ing biopolymers: polycaprolactone (PCL), polylactide Compliance with Ethical Standards 
(PLA) and thermoplastic starch (TPS). The homogeneous
morphology of TPS without the presence of accumulated Conflicts of Interest  The authors declare that they have no conflict of
starch granules has been achieved indicating successful interest.

13
Journal of Polymers and the Environment

References 33. Li R, Nie K, Shen X, Wang S (2007) Synth Charact Mater Lett
61:1368
34. Soares FC, Yamashita F, Müller CMO, Pires ATN (2014) Polym
1. Babu RP, O’Connor K, Seeram R (2013) Prog Biomater 2:8
Test 33:34
2. Fabra ME, López-Rubio A, Cabedo L, Lagaron JM (2016) J Col-
35. Silverajah VSG, Ibrahim NA, Zainuddin N, Yunus WMZW, Has-
loid Interface Sci 483:84
san HA (2012) Molecules 17:11729
3. Imre B, Pukánszky B (2013) Eur Poly J 49:1215
36. Janjarasskul T, Krochta JM (2010) Annu Rev Food Sci T 1:415
4. Song R, Murphy M, Li C, Ting K, Soo C, Zheng Z (2018) Drug
37. Mendes JF, Paschoalin RT, Carmona VB, Neto ARS, Marques
Des Devel Ther 12:3117
ACP, Marconcini JM, Mattoso LHC, Medeiros ES, Oliveira JE
5. Wachirahuttapong S, Thongpina C, Sombatsompop N (2016)
(2016) Carbohydr Polym 137:452
Energy Procedia 89:198
38. Kelnar I, Kratochvíl J, Kaprálková L, Zhigunov A, Nevoralová M
6. Fortelny I, Ujcic A, Fambri L, Slouf M (2019) Front Mater 6:206
(2017) J Mech Behav Biomed Mater 71:271
7. Sangeetha VH, Deka H, Varghese TO, Nayak SK (2016) Polym
39. Dziadek M, Stodolak-Zych E, Cholewa-Kowalska K (2017) Mater
Comp 39:81
Sci Eng C 71:1175
8. Siakeng R, Jawaid M, Ariffin H, Sapuan SM, Asim M, Saba N
40. Lee KH, Kimb HY, Khil MS, Ra YM, Lee DR (2003) Polymer
(2018) Polym Comp 40:446
44:1287
9. Correa AC, Carmona VB, Simão JA, Capparelli Mattoso LH,
41. Liao H-T, Wu C-S (2009) Mater Sci Eng A 515:207
Marconcinia JM (2017) Carbohydr Polym 167:177
42. Song Z, Huang X, Lu X, Lv Q, Xu N, Pang S, Pan L, Li T (2018)
10. Woodruff MA, Hutmacher DW (2010) Prog Polym Sci 35:1217
J Appl Pol Sci 135:1
11. Ferri Azor JM, Fenollar Gimeno OÁ, Jorda-Vilaplana A, García
43. Dias P, Chinelatto MA (2019) J Renew Mater 7:129
Sanoguera D, Balart Gimeno RA (2016) Polym Int 65:453
44. Persenaire O, Alexandre M, Degée P, Dubois P (2001) Biomac-
12. Urquijo J, Guerrica-Echevarría G, Eguiazábal JI (2015) J Appl
romol 2:288
Polym Sci 42641:1
45. Castro-Aguirre E, Iñiguez-Franco F, Samsudin H, Fang X, Auras
13. Navarro-Baena I, Sessini V, Dominici F, Torre L, Kenny JM,
R (2016) Adv Drug Deliv Rev 107:333
Peponi L (2016) Polym Degrad Stab 132:97
46. Ayumi SM, Müller CMO, Grossmann MVE, Yamashita F (2015)
14. Neto BAM, Fornari Junior CCM, Galvão E, da Silva P, Franco
J Polym Environ 23:54
M, Reis N-S, Cristina R, Bonomo F, de Almeida PF, Pontes KV
47. Sangeetha VH, Deka H, Varghese TO, Nayak SK (2016) Polym
(2017) Int J Food Prop 20:2429
Compos 39:81
15. Worzakowska M (2016) J Therm Anal Calorim 126:531
48. Aoyagi Y, Yamashita K, Doi Y (2002) Polym Degrad Stabil 76:53
16. Worzakowska M, Grochowicz M (2015) Carbohydr Polym
49. Carmona VB, Corrêa AC, Marconcini JM, Mattoso LHC (2015)
130:344
J Polym Environ 23:83
17. Khan B, Niazi MBK, Samin G, Jahan Z (2017) J Food Process
50. Yu L, Dean K, Li L (2006) Prog Polym Sci 31:576
Eng 40:1
51. Li J, Luo X, Lin X, Zhou Y (2013) Starch/Stärke 65:831
18. Pervaiz M, Oakley P, Sain M (2014) Int J Compos Mater 4:204
52. Feng CS, Piao M, Li D (2016) Polymers (Basel) 8:165
19. Long Y, Dean K, Li L (2006) Pro Polym Sci 31:576
53. Jia S, Yu D, Zhu Y, Wang Z, Chen L, Fu L (2017) Polymers 9:528
20. Mittal V, Akhtar T, Matsko N (2015) Macromol Mater Eng
54. Averousa L, Moroa L, Doleb P, Fringant C (2000) Polymer
300:423
41:4157
21. Zhang K, Mohanty AK, Misra M (2012) ACS Appl Mater Inter-
55. Avella M, Errico ME, Laurienzo P, Martuscelli E, Raimo M,
faces 4:3091
Rimedio R (2000) Polymer 41:3875
22. Parameswaranpillai J, Thomas S, Grohens Y (2015) In: Thomas
56. Xu X-R, Xu J-T, Feng L-X (2002) Polym Int 51:458
S, Grohens Y, Jyotishkumar P (eds) Characterization of polymer
57. Ostafinska A, Fortelný I, Hodan J, Krejčíková S, Nevoralová M,
blends: miscibility, morphology and interfaces, vol 1. Wiley-VCH,
Kredatusová J, Kruliš Z, Kotek J, Šlouf M (2017) J Mech Behav
Weinheim p, pp 1–6
Biomed Mater 69:229
23. Bulatovic OV, Grgić KD, Slouf M, Ostafinska A, Dybal J,
58. Jesús García-Campo MJ, Boronat T, Quiles-Carrillo L, Balart R,
Jozinović A (2019) Chem Pap 73:1121
Montanes N (2018) Polymers 10:3
24. Van Oss CJ, Giese RF, Li Z, Murphy K, Norris J, Chaudhury
59. Brostow W, Hagg Lobland HE, Narkis M (2006) J Mater Res
MK, Good R (1993) Contact angle wettability and adhesion. VSP,
21:2422
Utrecht
60. Brostow W, Hagg Lobland HE (2017) Materials: introduction and
25. Ahamed F, Phang SW, Sin LT (2016) JESTEC 11:1344
applications. Wiley, New York
26. Oduola MK, Akpeji PO (2015) Am J Chem Eng 3:1
61. Brostow W, Hagg Lobland HE, Hong HJ, Lohse S, Osmanson AT
27. Wu S (1973) J Adhesion 5:39
(2019) J Mater Sci Res 8:31
28. Le Corroller P, Favis BD (2011) Polymer 52:3827
29. Zolali AM, Favis BD (2017) Soft Mater 13:2844
Publisher’s Note Springer Nature remains neutral with regard to
30. Fu Y, Fodorean G, Navard P (2018) Polym Int 67:1378
jurisdictional claims in published maps and institutional affiliations.
31. Xing Z, Yang G (2010) J Appl Polym Sci 115:2747
32. de Campos A, Tonoli GHD, Marconcini JM, Mattoso LHC, Klam-
czynski A, Gregorski KS, Wood D, Williams T, Chiou B-S, Imam
SH (2013) J Polym Environ 21:1

13

You might also like