Sei sulla pagina 1di 12

J Polym Environ (2018) 26:330–341

DOI 10.1007/s10924-017-0951-3

ORIGINAL PAPER

Biodegradation in Soil of PLA/PBAT Blends Compatibilized


with Chain Extender
Paula Alessandra Palsikowski1 · Caroline N. Kuchnier1 · Ivanei F. Pinheiro1 ·
Ana Rita Morales1 

Published online: 7 February 2017


© Springer Science+Business Media New York 2017

Abstract  This paper presents a study of biodegradation, Introduction


in soil, of samples of poly(butylene adipate-co-terephtha-
late)(PBAT), poly(lactic acid) (PLA) and blends of these Increased production and consumption of plastic materials
materials prepared in torque rheometer with the addition has a direct effect on waste disposal causing a considerable
of a chain extender. Film samples of these materials were concern about the environmental impacts of the polymeric
buried in soil under controlled laboratory conditions. The materials, which has led researchers to search alternatives
degraded samples were regularly taken from soil and ana- in the biodegradable materials field [1–5]. Besides biode-
lyzed by visual inspection, size exclusion chromatography, gradable polymers presents potential for many different
differential scanning calorimetry and infrared spectros- applications including medical such tissue engineering,
copy. Respirometry biodegradation tests were conducted drugs delivery, orthopedic devices or packaging applica-
to assess samples mineralization degree. Blends showed tions. However most of these materials have some lack
higher degree of crystallinity compared to pure polymers. of properties that are object of study to be overcome by
Crystallinity degree enhanced during the biodegradation modification as blends or composites [6]. One of the most
process in all samples, being able to causing the samples to studied material is the poly(lactic acid) (PLA), which is an
degrade slowly. The study showed the great complexity of aliphatic polyester that can be synthesized from renewable
the biodegradation process of PLA and PBAT blends when resources. PLA is brittle and presents low susceptibility to
compatibilized with a chain extender. The biodegradation microbial attack on natural environment. Its degradation is
rate showed different results depending on the characteris- highly dependent on environmental conditions due to its
tic applied to evaluate it: visual, molecular weight or miner- high glass transition temperature (­ Tg) [7, 8]. The T
­ g of PLA
alization. The chain extender had strong influence in PBAT is above room temperature, which limits composting condi-
and blends degradation, slowing the process as observed tions biodegradation, therefore soil degradation is very low
by the variation of molecular weight and carbonyl index. [1, 7].
Blends showed an intermediate behavior compared to the PLA has been widely used in packaging, nevertheless its
original polymers. application could be extended with mechanical properties
optimization aiming to brittleness reduction and improvement
Keywords  Poly(lactic acid) · Poly(butylene adipate- of soil degradation susceptibility. Blends with poly(butylene
co-terephthalate) · Biodegradation · Soil biodegradation · adipate-co-terephthalate) (PBAT) have been extensively
Compatibilized blends studied with purpose of producing materials with both PLA
and PBAT advantageous properties [2, 8–18]. PBAT is an
aliphatic–aromatic biodegradable polyester produced from
* Ana Rita Morales fossil sources [19]. The aromatic fraction provides excel-
morales@feq.unicamp.br lent physical properties whereas aliphatic chains promote its
1
School of Chemical Engineering, State University
degradation in several conditions, including soil degradation
of Campinas (UNICAMP), Av. Albert Einstein 500, without the need of temperature control [19, 20]. PBAT can
Campinas, SP 13083‑852, Brazil modify PLA fragile properties, although it was reported to be

13
Vol:.(1234567890)
J Polym Environ (2018) 26:330–341 331

incompatible with PLA [11]. Many studies aimed at improv- PLA Ingeo 4042D, density 1.24  g/cm3, average molec-
ing compatibility between these materials, by functional ular weight (Mw ) 201,000  g/mol and number-average
groups chemical reaction like in-situ compatibilization using molecular weight (Mn) 101,000 g/mol from Natureworks;
dicumyl peroxide as a free-radical initiator [2, 21, 22] or PBAT Ecoflex® F BX 7011, density 1.25  g/cm3, Mw
using a chain extender with multifunctional groups that can 105,000 g/mol and Mn 47,000 g/mol, from BASF and ;
react with both carboxyl and hydroxyl groups during the pro- Multifunctional chain extender Joncryl ® ADR-4368,
cessing of PLA and PBAT [9–11, 16, 18, 23, 24]. Multifunc- density 1.08 g/mL, Mw 6,800 g/mol and equivalent weight
tional chain extenders are thermally stable substances that of functional epoxy groups of 285 g/mol, from BASF.
promote increase in molecular weight by generation of long
linear or branched chains. Their action is dependent on poly-
mers chains scission that occurs during processing, which Sample Preparation
generates reactive groups [25]. The most common chain
extender, Joncryl® ADR-4368, has epoxy groups capable of The polymers pellets were dried in an oven at constant
reacting with the OH and COOH and promote the recovery temperature (PLA 80 °C for 4  h and PBAT 70 °C for 40
of recycled polymer properties like poly(ethylene terefhtha- min). Samples were blended in the proportions: 100%PLA
late) (PET) [25–28]. (PLA), 75%PLA/25%PBAT (75/25), 25%PLA/75%PBAT
PLA degradation has been widely studied as pristine (25/75) and 100%PBAT (PBAT), all added with 1 part of
polymer [1, 7, 29–31] and, recently, as modified polymer chain extender per hundred parts of resin (1 phr) and pro-
[19, 32–38]. The mechanical properties of PLA/PBAT cessed in a torque rheometer HAAKE Rheomix 600P with
blends were reported in many works were the brittle prop- a rotation speed of 120 rpm for 5 min at 180 °C. Samples
erties of PLA could be modified by the addition of PBAT were pressed in a hydraulic press LP20-B from Labtech
probably related to the presence of a soft elastomeric phase Engineering Co. Ltd. The molten polymers were molded
and the compatibilization effect of the Joncryl ® ADR- at constant pressure of 800 psi at 120 °C (PBAT), 160 °C
4368 [39]. The elongation at break of the PLA/PBAT (PLA) and 150 °C (blends) for 1 min. Then, samples were
blends was reported to increased with the incorporation cooled at room temperature, obtaining films with thick-
of Joncryl® ADR-4368 more than 100% which was asso- ness of approximately 0.3 mm. All conditions were applied
ciated to formation of ester linkages between PLA, PBAT, based on a previous work [43].
and Joncryl® ADR-4368 at the interface of the polymers.
Blown films prepared from similar systems showed a sig-
nificant improvement in the ductility as a combination of Evaluation of Soil Biodegradability
interfacial tension and blends viscosity ratio that resulted
in a finer and homogeneous morphology [10]. The studies Although the most usual studies for the biodegradation
show that amounts from 0.25 up to 1% of Joncryl® ADR- for PLA are done under composting conditions, our work
4368 are enough to promote improvement of general prop- aim to evaluate the behavior of the blends in soil, since
erties. Thus the addition close to 1% of chain extender Jon- the PBAT was supposed to bring to the system the char-
cryl® ADR-4368 was used in this work, as it has shown acteristic of high disintegration rate in soil conditions. The
an improvement on the thermal stability and mechanical biodegradation process of pure polymers and blends was
properties of PLA and PBAT [11, 42]. Even though PLA/ analyzed in three ways: visual analysis, molecular weight
PBAT blends have been, extensively studied, lacking has monitoring and carbon mineralization.
been reported about soil biodegradation and those studies
were limited for non-compatibilized blends [41, 42], or car-
ried out in alkaline solution [23]. Soil Preparation
The complexity of the biodegradation process motivated
this work, which objective was the study of the biodegra- The soil preparation followed the ASTM D5988-12,
dation process of PLA/PBAT blends compatibilized with a which is the standard for determining aerobic biodegrada-
chain extender in the soil under laboratory conditions. tion of plastic materials in soil. A mixture of soil aliquots
collected into three different regions was used to ensure
microbial biodiversity. The mixed soil was dried for 72 h at
Materials and Methods room temperature (25  ±  2 °C) and then sieved in order to
obtain particle sizes smaller than 2 mm. The soil was char-
Materials acterized and adjusted to fit the tests according to specific
methodologies related with their characteristics shown in
The materials used were: Table 1.

13

332 J Polym Environ (2018) 26:330–341

Visual Analysis and Monitoring of Molecular Weight was constructed with polystyrene with molecular weight in
a range of 1.1 kDa and 3.8 MDa.
Samples were degraded as described by ASTM D5988-
12. Soil moisture was adjusted and maintained at 60% of Respirometry Test
its retention capacity, maximum amount of water that the
soil can retain after the excess has been drained, accord- The respirometry test was performed according to ASTM
ing ASTM D5988-12. Eight square samples of 3  cm x D5988-12 standard adapted according to the methodology
3 cm and thickness around 0.3 mm were subjected to soil proposed by Anderson [44]. Films were subjected to ele-
degradation in a closed container maintained at room tem- mental analysis to determine carbon amount (mg/100  mg
perature (25  ±  2 °C). The container base was covered with sample) from the stoichiometry of carbon reaction (C)
soil (approximately 1.5  cm); samples of each material (Eq. 1) 12 g of C produced 44 g of carbon dioxide ­(CO2).
were placed on this layer being subsequently covered with
approximately 6 cm of soil. Every 30 days, one sample of
C + O2 → CO2 (1)
each material was taken for visual appearance evaluation Thus:
and buried again.
44 ∗ Y
During this period, one fragment of each material was MtCO2 (mg) = (2)
removed for monitoring of the molecular weight by Size 12
Exclusion Chromatography (SEC). In this analysis, the where MtCO2: ­CO2 theoretical mass. Y is the amount of C
samples were solubilized in THF stabilized with BHT (mg) in the sample.
(2,6-Di-tert-butyl-4-methylphenol), concentration of The actual amount of ­CO2 produced (mg) is given by
250 ppm, 99.9%, Scharlau, HPLC degree with 5 mg/mL of Eq. (3).
concentration. It was used three columns in series (Shodex,
KF-806M); all are filled with a copolymer of styrene and
MpCO2 (mg) = (Vcn − Vam) ∗ 22 ∗ 0.5 (3)
vinyl benzene, with exclusion band from ­103 to ­106. Ana- where MpCO2: mass of ­CO2 produced (mg). Vcn is the
lyzes were performed by a chromatograph Viscotek model volume of HCl 0.5  N used in the titration of the negative
GPC Max, using refractive index detector model VE3580 control flask; Vam is the volume of HCl 0.5 N used in the
at 40 °C under a flow of 1 mL/min. The calibration curve titration of the sample flask; 22: Equivalent gram of C ­ O2;
0.5: Normality of HCl.
Portions of 200 g of dried soil were placed in glass flasks
with sealing caps. The films, in duplicate, were weighed
Table 1  Soil characteristics and buried in the soil portions. Two flasks containing only
Tests Parametermethodology Results soil as blank samples and other two flasks containing soil
and cellulose were used as positive control. Deionized
Visual analysis, monitor- Moisture (%)a 41 water was added to each flask so that the moisture was 60%
ing of molecular weight Retention capacity (HC) (%)a 75 of the soil retention capacity (36% moisture). Into each
Moisture test (%) 45 flask a beaker with 20.0 mL of NaOH 0.5 N was placed to
pHb 7.3 capture the C ­ O2 and a beaker with 50 mL distilled water
Organic matter (g/dm3)b 72 was placed in order to keep the moisture. It is important
Organic carbon (g/dm3)3 40 to mention that the presence of the water does not inter-
Nitrogen (g/kg)b 2.7 fere in the capture of the ­CO2, since the reaction between
C/N Ratio 14.8 ­CO2 and H ­ 2O generates H­ 2CO3 which converts again to
Mineralization test Moisture (%)a 4.7 ­CO2 and ­H2O. The flasks were opened, the beaker with
Retention capacity (HC) (%)a 60 NaOH was removed, and 1.0  mL of saturated solution of
Moisture test (%) 36 barium chloride ­(BaCl2) was added to cause the carbonate
pHb 6.3 precipitation. The solution of NaOH 0.5 N was then titrated
Organic matter (g/dm3)b 45 with HCl 0.5 N in order to obtain the ­CO2 captured by the
Organic carbon (g/dm3)c 25 NaOH solution. After titration, a beaker with the same vol-
Nitrogen (g/kg)c 1.5 ume of NaOH 0.5 N freshly prepared was added for another
C/N Ratio 16.7 degradation period. Beakers containing distilled water were
a
 Casado (2009) replaced by fresh distilled water at every titration. In the
b
 methodologies described by the Ministry of Agriculture, Livestock first 30 days, these evaluations were made twice a week,
and Supply (MAPA, 2007) and then once a week for the last 90 days. This experiment
c
 Kiehl (1985) was performed in duplicate.

13
J Polym Environ (2018) 26:330–341 333

The mineralization (amount of polymer C converted Ic


in ­CO2) of the formulations was calculated according to %C = × 100 (5)
Ic + Ia
Eq. (4).
where %C is the crystalline fraction. Ic is the diffraction
MpCO2
Mineralization (%) = (4) peaks area. Ia is the amorphous halo area.
MtCo2
Fityk 0.8.0, an open-source curve-fitting and data anal-
where ­MpCO2CO2 mass produced (mg), ­ MtCO2 = Total ysis software (distributed under the terms of GNU Gen-
theoretical ­CO2 mass (mg). eral Public License), was used to obtain the area of the
Mineralization tests were performed with PLA, PBAT amorphous halo and of the crystalline peaks.
and the 25/75 and 75/25 blends, all with chain extender
addition. To evaluate the soil activity, ASTM D5988-12
recommends the use of a biodegradable material as refer-
ence to evaluate soil activity; cellulose was used in this Results and Discussion
work. The average amount of released C ­ O2 was monitored
as a function of time, and two replicates were performed Visual Analysis
for each material.
Changes in samples appearance related to different degra-
Samples Characterization dation times are shown in Fig. 1. Samples were gradually
degrading over time when buried in soil. Biodegradation
Fourier Transform Infrared Spectrometer Analysis (FTIR) under laboratory conditions are not as aggressive as the
actual conditions, especially in terms of temperature. Even
Identification of functional groups in the chemical struc- in milder cases, microbiological attack on samples was
tures of the materials was carried out by Fourier Transform apparent, presenting yellowing and dark brown spots on
Infrared Spectrometer Analysis (FTIR), using a Thermo the surface of the films. Although the non-compatibilized
Nicolet spectrometer, model Nexus 470, Thermo Scientific. blends are not interesting for not presenting satisfactory
One sample of each material was analyzed from 4000 to mechanical properties, a shorter assessment of these sys-
675 cm−1 with resolution of 4 cm−1. These analyzes were tems was chosen, 30 days, in order to understand the partial
performed before and after 360 days of degradation. results, especially related to PBAT with chain extender that
did not show any signs of degradation. It was observed that
Differential Scanning Calorimetry Analysis (DSC) pristine PBAT became very fragile and brittle, leaving only
fragments after 30 days of degradation, preventing continu-
To evaluate the thermal transitions of samples and the ity of visual and molecular weight analysis. Changes were
degree of crystallinity, Differential Scanning Calorimetry not visually observed in samples of PLA and 75/25 blend
Analysis (DSC) was performed on a DSC-2920 Modulated after 30 days of degradation, whereas 25/75 blend had some
DCS equipment, TA Instruments, under N ­ 2(g) flow of 100 yellow spots. For materials modified with chain extender
mL/min. According to ASTM D3418-12 standard, the steps we observed, after 120 days, minor yellow spots on the
performed were: heating from room temperature to 200 °C, PBAT film surface, which had the first break after 180
10 °C/min, isotherm for 5 min at 200 °C; cooling from 200 days; after 270 days it was already too brittle and after 360
to −80 °C, 10 °C/min, isotherm for 5 min at −80 °C; heat- days only small dark fragments were left. The PLA sample
ing from −80 to 200 °C, 10 °C/min. As biodegradation ini- presented cracks after 270 days and breaks after 360 days.
tiates, changes in the material crystallinity can occur, so By visual analyses, PLA showed to be much less degrada-
these analyzes were performed before and after 360 days of ble than samples containing PBAT. The higher resistance
degradation, one sample of each material. to biodegradation of PLA in soil has been already reported
and is assumed to be due to its glass transition temperature
X‑Ray Diffraction (XRD) ­(Tg), around 60 °C [1]. Blend samples showed intermediate
behavior compared to pure materials. First, blends showed
Crystallinity was also analyzed by X-Ray Diffraction yellowing followed by some deformations, and then breaks
(XRD) using a diffractometer model X’Pert-MPD, Philips after 180 days. The blend with higher amount of PBAT
Analytical X Ray, scanning range from 5° to 35° (2ϴ), Ka was less resistant to degradation. These results indicate the
radiation of Cu (λ  = 1.54  Å), 40  kV, 40  mA, 0.02° step, influence of chain extender, delaying the PBAT degradation
0.02º/s, at room temperature. Crystallinity was calculated process that when pure, presented very high degradation
as the ratio between the crystalline area (peaks) and amor- rate, consistent with other studies [42].
phous halo, by Eq. (5):

13

334 J Polym Environ (2018) 26:330–341

Fig. 1  Visual aspect of the


samples during degradation in
soil

Monitoring of Molecular Weight expected since the epoxide groups of the extender are very
reactive with the carboxyl and hydroxyl groups of PLA,
SEC analysis were performed before and after 60, 120 and increasing the size of chains. However, Joncryl® ADR-
180, 240, 300 and 360 days of degradation, for samples 4368 was less effective in the blends and did not alters to
containing chain extender and after 30 days of degrada- much the PBAT molecular weight, a result also reported
tion for samples without chain extender. Results for number [43]. However, comparing Mw of PBAT with and without
average molecular weight (Mn), weight average molecu- the chain extender, it is noted that the chain extender influ-
lar weight (Mw) and polydispersity (PDI) are shown in enced PBAT degradation process.
Table 2. For a better understanding of the biodegradation behav-
The hydrolysis process was evaluated considering Mw ior, we can see in Fig. 2b the Mw on percentage as a func-
because it is less affected by low molecular weight than tion of degradation time in which it is evident the influ-
Mn [23]. Figure  2 shows the Mw results. Generally, there ence of chain extender on the materials degradation. It was
was a decrease in average molecular weight over the deg- observed that there was initially a small reduction on PBAT
radation time indicating the occurrence of chain scission. Mw and that after 180 days the Mw increased. Such behav-
Mw values of samples without chain extender are also pre- ior indicates that the epoxide groups of the chain extender
sented. Initially (t = 0 days) it was observed that the molec- are reacting with the groups that are originated during
ular weight of PLA was much smaller than the molecu- PBAT degradation process, increasing Mw. Although
lar weight of PLA with Joncryl® ADR-4368, which was PBAT degradation visually occurs, the results indicate a

Table 2  Data obtained by Sample Days 0 60 120 180 240 300 360 Sample 0 30
SEC analysis for different
degradation times PLA w/Joncryl® Mn—(KDa) 126 147 149 149 107 105 109 PLA 64 62
Mw—(KDa) 449 349 312 312 239 239 244 113 107
PDI 3.6 2.4 2.1 2.1 2.2 2.3 2.2 1.8 1.7
75/25 w/Joncryl® Mn—(KDa) 36 35 23 33 23 20 20 75/25 17 13
Mw—(KDa) 196 148 142 139 151 129 133 91 82
PDI 5.5 4.2 6.1 4.2 6.7 6.3 6.8 5.4 6.2
25/75 w/Joncryl® Mn—(KDa) 21 20 21 21 14 14 13 25/75 13 13
Mw—(KDa) 86 79 84 74 58 63 53 45 33
PDI 4.2 3.8 4.0 3.6 4.2 4.5 4.1 3.5 2.6
PBAT w/Joncryl® Mn—(KDa) 19 17 17 18 13 13 11 PBAT 23 18
Mw—(KDa) 44 43 45 42 56 46 29 51 29
PDI 2.3 2.5 2.6 2.3 4.4 3.5 2.5 2.2 2.1

13
J Polym Environ (2018) 26:330–341 335

Fig.  2  a Weight average molecular weight (Mw) and b Relative molecular weight (Mw) as a function of degradation time

competitive effect: chains scission reduces the molecular chain segments are consumed by microorganisms in soil,
weight and creates functional groups, which react with the not being measured by SEC analysis. Although the chain
chain extender, increasing the molecular weight and slow- extender has initially increased PLA molecular weight, it
ing the process. did not slow the PLA degradation. The degradation behav-
A degradation parameter k can be considered to evaluate ior of 25/75 blend with chain extender was very similar to
the chain scission as described by the Eq. (6) [7], where Mn the PBAT with chain extender until the 120th day of deg-
BD refers to the material before degradation and MnAD after radation, both in terms of Mw and visual characteristics.
degradation. The results were shown in Table 3. Continuing the degradation process, the blend showed
behavior more similar to PLA with chain extender, with
MnBD Mw reduced at a greater rate between the 120th and the
k= (6) 240th day of degradation. This result may be associated
MnAD
with material morphology. According to a previous study
Signori et  al. [8] reported the high stability of PBAT [43] the 25/75 blend presents a PLA dispersed phase in the
upon processing. This behavior was also observed by Al- PBAT matrix. The PBAT surfaces are available for the ini-
Itry et  al. for the effect of the processing temperature on tial degradation and the blend behavior is similar to PBAT
the thermal stability for the same system studied here [11]. degradation. Over degradation time, PBAT chains are bro-
They showed k = 1.01 for PBAT processed at 180 °C. Using ken and released to the medium and/or consumed by the
the k factor for the degradation process we can observe microorganisms, leaving the PLA rich phase more exposed
k = 1.28 for 30 days of buring. It means that the chains scis- to the soil showing a behavior similar to PLA degradation.
sions are very pronounced during the biodegradation even The chain extender influenced the degradation of this blend
at room temperature. As the Joncryl® acts on the OH and in the same way as PBAT. On the other hand, 75/25 blend
COOH groups resulted of the biodegradation process, there with chain extender showed initial degradation behavior
will be much more available points to react than during similar to PLA with chain extender addition; however, after
the processing as can see by the k value of 1.12 for sam- 180 days this blend had an increase in Mw, most likely due
ples with Joncryl® after 60 days of buring. The difference to chain extender action in PBAT phase. The initial deg-
between SEC and visual results may be due to mainly deg- radation behavior in the blends was similar to the matrix;
radation of PBAT surface by the action of enzymes. Small when the surface of the domains started to be exposed due

Table 3  K factor for the Sample /days 0 60 120 180 240 300 360 0 30
degradation process calculated
from the GPC data PLA w/Joncryl ® 1.00 0.86 0.85 0.85 1.18 1.20 1.15 PLA 1 1.03
75/25 w/Joncryl ® 1.00 1.02 1.57 1.09 1.58 1.77 1.83 75/25 1 1.29
25/75 w/Joncryl ® 1.00 1.01 0.98 0.99 1.49 1.45 1.62 25/75 1 1.00
PBAT w/Joncryl ® 1.00 1.10 1.10 1.03 1.49 1.46 1.67 PBAT 1 2.70

13

336 J Polym Environ (2018) 26:330–341

to degradation, the behavior of the dispersed phase mate- long chains that is required for subsequent mineralization
rial was observed. Other studies on the chain extender [46]. After 20 days of mineralization PLA reached less
effect and the PBAT presence in the PLA degradation in than 2% of mineralization. PLA embrittlement occurs when
alkaline solution at 60 °C showed that PLA hydrolysis was the molecular weight reaches a value close to 10–20 kDa,
little affected [23], differently from what was observed in which is a critical value for microorganisms continue the
this study, indicating that the degradation process of these degradation process, converting low molecular weight
blends is very sensitive to the adopted conditions. components into metabolites [49, 55, 56]. Table  2 shows
that the average PLA molecular weight, even after 360 days
Mineralization of degradation, is clearly above this value; however, dur-
ing the degradation few PLA chains on the material surface
Figure  3 shows the results of the biodegradation test by can be broken releasing small segments to be consumed by
mineralization of carbon. The amount of ­ CO2 released microorganisms.
from the cellulose (positive control) during biodegradation PBAT had the highest degree of mineralization, reach-
increased rapidly in the first 20 days at a rate of 2.72 mg/ ing approximately 9% after 60 days of degradation and
day, reaching approximately 20% mineralization in 30 days. 21% after 180 days, comparable with other studies [46].
After that, the mineralization rate decreased. It can be seen The mineralization rate was 0.23  mg/day, being greater
that cellulose has reached about 52% mineralization within in the first 60 days of testing. PBAT undergoes hydrolytic
180 days, which, according to ASTM D5988-12 standard, and microbiological degradation due to presence of heter-
does not validate the test, since 70% of the carbon mate- oatoms, carbonyl and aliphatic chain, being more suscep-
rial should be converted into ­CO2 after 180 test days. Since tible for biodegradation [3, 7]. Its degradation is acceler-
microbiological activity in soil interferes with the degrada- ated by the presence of enzymes that are released by the
tion rate [45], this result indicates that the soil could have microorganisms in soil. Considering the results of SEC
low amounts of decomposing microorganisms. This is con- analysis, visual analysis and mineralization test we can say
sistent with previous studies of PLA and PBAT having cel- that the PBAT degradation occurs preferably on the surface
lulose as positive control in two types of soil, a sterilized of the material, which is primarily caused by the action
and an inoculated with fungi; the latter had high biological of enzymes, releasing low molecular weight chains to the
activity [46]. environment, which are consumed and converted into ­CO2
The mineralization rate of carbon for PLA was about and are not measured by SEC analysis.
0.17  mg/day, reaching a percentage of mineralization of The blends showed lower mineralization degree val-
approximately 16% in 180 days; results comparable to ues to those obtained for pristine materials, with rates of
those were reported a mineralization rate of 0.24  mg/day 0.15 mg/day for the 25/75 blend, and 0.12 mg/day for the
and mineralization of 17% after 182 days [47]. In the first 75/25 blend.
days of the experiment, there was a very little release of
­CO2. This delay corresponds to the hydrolysis step of the
Infrared Spectroscopy (FTIR)

Table 4 shows the bands in the FTIR spectra of the samples


before and after 360 days of degradation in order to evalu-
ate what modifications occurred on the samples surface. No
differences before and after degradation were observed on
the bands, as the wave number changes are in the resolution
range of the equipment (4  cm−1). Rudnik and Briassoulis
[7] measured the intensity of the absorbance at 1748 cm−1
corresponding to carbonyl stretching band to monitorate
the degradation of PLA buried in soil. They found values
changing less than 1 from 3.24 to 2.44 and from 3.6 to 3.5
depending on the PLA grade. Tham et  al. [48], found an
increase of more than 2 of the CI after hydrothermal aging
and associated these increase to the hydrolysis of PLA.
From the spectra of our samples, the carbonyl index (CI)
was calculated as the ratio between the area of the carbonyl
Fig. 3  Mineralization test of material tests and cellulose as the peak and the area of the band of -CH- peak, chosen as
ASTM D5988-12 at 28° C reference.

13
J Polym Environ (2018) 26:330–341 337

Table 4  Infrared spectral PBAT band position ­cm−1 Position of band 25/75 Position of band 75/25 PLA band position
analysis of the samples before ­cm−1 ­cm−1 ­cm−1
and after degradation
0 days 360 days 0 days 360 days 0 days 360 days 0 days 360 days

2998 2996
2957 2957 2956 2960 2958 2956 2946 2946
1756 1756 1754 1754
1712 1712 1712 1712 1712 1714
1475 1475
1455 1455 1458 1458 1456 1452 1454 1452
1409 1409 1409 1409 1409 1396 1361 1365
1387 1387 1390 1363
1359 1358
1271 1271 1272 1270 1270 1269 1268 1268
1221
1162 1166 1162 1168 1182 1184 1184 1184
10
1105 1105 1105 1105 1103 89 1132 1132
1087 1087
1045 1045 1045 1045
1018 1018 1018 1018 1018
932 937 933 935 931 917
873 873 873 873 873 871 866 867
727 727 727 727 729 730 750 750

depends of the environment where the degradation occurs.


It was observed discreet CI variations and no IR spectra
alterations after degradation. These results indicate that the
degradation process was slow and that there was no forma-
tion of byproducts on the samples surface [6]. Moreover,
slight changes corroborate the chain extender action during
the degradation process.

Analysis of Differential Scanning Calorimetry (DSC)

Figure 5 shows the curves of the first heating on the DSC.


The data regarding Tg and Tm in Table 5 were taken from
second heating.
Figure  5 shows enthalpy relaxation for PLA that has
been observed in other studies [36, 49, 50] and typically
occurs for polymers with glass transition temperature near
Fig. 4  Carbonyl index of the samples before and after 360 days of room temperature. An exothermic peak can be seen related
degradation
to cold crystallization for pristine PLA, for samples before
and after 360 days of degradation, for 75/25 blend before
The results are shown in the Fig.  4. For PLA, CI was degradation and it was not observed for 25/75 blend. This
calculated for the band at 1754  cm−1 (C = O) and at phenomenon occurs when polymer molecules, under heat-
1452 cm−1 (reference) and changed from 4.94 to 5.13. For ing, acquire sufficient mobility to rearrange its crystal
the PBAT, bands at 1712 and 1455 cm−1 were considered, structure. The melting of PBAT is shown as a wide tem-
and CI changed from 2.96 to 4.54. The CI of the 25/75 perature range and is very close to PLA cold crystalliza-
blend changed from 3.06 to 3.48 and for the 75/25 blend, tion such that these effects were overlapped in blends. This
the CI changed from 4.63 to 3.09. These different results fact hinders the analysis of PBAT melting and PLA cold
suggest that the degradation is quite complex and that crystallization effect in blends, leading to inaccurate values

13

338 J Polym Environ (2018) 26:330–341

PLA cold crystallization, which may become in crystalline


form ordered α [50, 51].
The curves for the second heating are shown in Fig. 6.
There was not large variation before and after 360 days
of degradation. Identifying the PBAT fusion peak in the
75/25 blend was not possible neither was the PLA cold
crystallization in the 25/75 blend due to the superposition
of transitions. Also the Tg of PLA in the blends had no
important variation, which was expected since the materi-
als are immiscible, although a decreasing trend could be
related to a little degree of miscibility caused by the chain
extender. The Tg of PBAT in the blends was only observed
in 25/75 blend, since the lowest PBAT concentration in the
75/25 blend did not allow the measurement by the limited
DSC sensitivity. However, after degradation, both PLA
and PBAT undergo a slight increase in their Tg. The larg-
Fig. 5  Curves of the first heating of DSC analysis. a PLA—0 days;
b PLA—360 days; c 75/25—0 days; d 75/25—360 days; e 25/75—0 est segments of chains of the amorphous phase are less
days; f 25/75—360 days; g PBAT—0 days; h PBAT—360 days restricted by the crystalline phase (lower Tg) and are easily
broken being bio-assimilated or crystallized. The remain-
ing segments in the amorphous phase are smaller and are
in calculation of the PLA crystallinity. Due to this super- restricted by the crystalline phase, presenting greater diffi-
position, XRD analyzes were performed to obtain the total cult to degrade resulting in the highest Tg values.
degree of crystallinity of the blends (%C), which is also Crystallinity degree was calculated to assess the effect of
shown in Table 4. degradation in the crystallization as Eq. (7).
In the first heating curve a shoulder preceding the melt-
ing peak of the 75/25 blend was observed, which may be ΔHf − ΔHcc
related to the existence of two crystalline phases in this
Xc1 =
w.ΔHf∞
× 100 (7)
blend. PBAT may have induced the formation of a new
crystalline structure in PLA [12]. According to Fukushima where ΔHf is the heat of fusion; ΔHcc is the enthalpy of
et  al., (2009) the melting crystalline phase at lower tem- cold crystallization of PLA; ΔHf∞ is the enthalpy of fusion
perature is called phase α, whereas the higher temperature for a crystal with infinite thickness (93 J.g− 1 for PLA and
melting peak is characteristic of crystals of type. The heli- 114 J.g− 1 for PBAT) [11]; w is the PLA mass fraction.
cal conformations of the PLA chains in structures α and β Xc1was calculated from the first heating curves as repre-
have approximately the same energy, it is believed that the senting the conditions of the specimens and Xc2 from the
way each of them is organized is the main reason for the second heating as the characteristic of the materials, where
existence of two different crystalline structures [58]. Also ΔHcc has not been subtracted, since the objective was to
a disordered crystalline phase α’ has been described during calculate any crystallinity present in the material.

Table 5  Thermal properties of Sample-days of Tg (oC) Tcc (oC) Tm (oC) Xc1(%) Xc2(%) %C (%)
the samples before and after 360 degradation
days of degradation PLA PBAT PLA PLA PBAT

PLA—0 60 – 115 148 – 0.9 22.2 0


PLA—360 61 – 116 150 – 1.3 25.7 0
75/25—0 59 – 126 150 – 13.1* 11.1* 35.4
75/25—360 60 – 128 152 – 40.6* 13.9* 41.3
25/75—0 58 −34 – – 129 11.3* 27.7* 28.2
25/75—360 60 −31 – – 132 20.2* 42.6* 30.2
PBAT—0 – −32 – – 130 15.2 8.4 17.8
PBAT—360 – −31 – – 132 15.3 12.2 26.5

*Regarding the PLA crystallization in the blend


%C crystallinity degree from XRD, Xc1 crystallinity degree from DSC 1st heating; Xc2 crystallinity degree
from DSC 2nd heating;

13
J Polym Environ (2018) 26:330–341 339

The higher degree of crystallinity of the blends before


degradation shows that the PBAT influenced PLA
crystallization.
All samples presented an increase in the degree of crys-
tallinity after the degradation. The increase of Xc1 indicates
that PLA hydrolysis occurs at a greater rate in the amor-
phous region that reduces the size of chain segments in this
region, which acquire mobility increasing the crystallinity
of the material as degradation occurs [30]. The increase of
Xc2,after degradation, is one more evidence that the crys-
tallization of the PLA is favored by the reduction of the
molecular weight and the presence of PBAT.

X‑ray Powder Diffraction (XRD)

One XRD analysis was performed for each material in


Fig. 6  Curves of the second heating of DSC analysis. a PLA—0 order to monitor changes in the structural order of the poly-
days; b PLA—360 days; c 75/25—0 days; d 75/25—360 days; e
25/75—0 days; f 25/75—360 days; g PBAT—0 days; h PBAT—360
mers due to degradation. Figure 7 shows the XRD patterns
days of samples before and after 360 days of degradation, all

Fig. 7  XRD patterns of a PLA, b 75/25 blend, c 25/75 blend and d PBAT before and after 360 days of degradation

13

340 J Polym Environ (2018) 26:330–341

with the addition of the chain extender. The degree of crys- Acknowledgements  The authors would like to thank the Foun-
tallinity (%C) is shown in Table 5. dation for Research of the State of São Paulo—FAPESP (Process
2014/09883-5) and National Council of Scientific and Technological
In PLA XRD patterns a broad peak at 2θ = 16° is Development CNPQ for the financial support, and the MICROMAT
observed, indicating a predominantly amorphous structure. (Prof. Lucia Innocentini Mei) for the mineralization tests.
These results are consistent with the DSC, which showed
a crystallinity degree around 1%, as it was very small it
was not detectable by the XRD analysis. PBAT XRD pat- References
terns before and after degradation show characteristics of a
semi-crystalline structure, with five well-defined diffraction 1. Rudeekit Y, Numnoi J, Tajan M, Chaiwutthinan P, Leejarkpai T
peaks at about 16.1°, 17.5°, 20.6°, 23.2° and 25.0°. All the (2008) JOM J Min Met Mat S 18:83–87.
2. Sirisinha K, Somboon W (2012) J Appl Polym Sci
XRD patterns of the 25/75 blend (before and after degrada- 124:4986–4992
tion) showed diffraction peaks similar to PBAT. The 75/25 3. Fukushima K, Rasyida A, Yang MC (2013) Appl Clay Sci
blend, different from the pristine polymers, presented a 80–81:291–298
peak at 2θ = 16.9° before degradation and 2θ = 16.6° after 4. Hughes J, Thomas R, Byun Y, Whiteside S (2012) Carbohyd
Polym 88:165–172
of degradation and low intensity peaks at 14.8°, 19° and 5. Livi S, Bugatti V, Marechal M, Soares BG, Barra GMO, Duchet-
22.3°. This data has been reported in other studies assigned Rumeau J, Gérard JF (2015) RSC Adv 5:1989–1998.
to the crystalline phase of PLA [52, 53]. The presence of 6. Hamada K, Kaseema M, Koa YG, Derib F (2014) Polym Sci Ser
these diffraction peaks indicate that in this composition 56:812–829
7. Rudnik E, Briassoulis D (2011) Ind Crop Prod 33:648–658.
PBAT influenced PLA crystalline structure, reinforcing the 8. Signori F, Coltelli MB, Bronco S (2009) Polym Degrad Stab
behavior observed in the curves in the first DSC heating 94:74–82
where 75/25 blend showed a bimodal endothermic peak. 9. Kumar M, Mohanty S, Nayak SK, Rahail Parvaiz M (2010)
By the XRD technique, except for the PLA, all samples Bioresour Technol 101:8406–8410
10. Arruda LC, Magaton M, R. E. S. Bretas, Ueki MM (2015)

had an increased degree of crystallinity due the degrada- Polym. Test 43:27–37
tion. This increase in crystallinity is one of the evidences 11. Al-Itry R, Lamnawar K, Maazouz A (2012) Polym Degrad Stab
of the biodegradation process [29, 30, 33, 54]. There is a 97:1898–1914
difference between the results obtained by the DSC and 12. Jiang L, Wolcott MP, Zhang J (2006) Biomacromolecules

7:199–207
XRD techniques. However, for pure polymers, these results 13. Lin S, Guo W, Chen C, Ma J, Wang B (2012) Mater Des

are quite consistent, and for the blends it is necessary to 36:604–608
remember that the calculated value corresponds to crystal- 14. Pivsa-Art W, Chaiyasat A, Pivsa-Art S, Yamane H, Ohara H
linity of both materials in the blends, while in relation to (2013) Energy Procedia 34:549–554.
15. Yeh JT, Tsou CH, Huang CY, Chen KN, Wu CS, Chai WL
DSC, PLA crystalline fraction was not possible due to the (2010) J Appl Polym Sci 116:680–687
superposition effect of transitions already reported. 16. Al-Itry R, Lamnawar K, Maazouz A, Billon N, Combeaud C
(2015) Eur Polym J 68:288–301
17. Touchaleaume F, Martin-Closas L, Angellier-Coussy H, Chevil-
lard A (2016) Chemosphere 144:433–439
Final Remarks 18. Wang LF, Rhim JW, Hong SI (2016) Food Sci Technol Int

68:454–461.
The study showed the great complexity of the biodegrada- 19. Fukushima K, Abbate C, Tabuani D, Gennari M, Camino G
tion process of PLA and PBAT blends when compatibilized (2009) Polym Degrad Stab 94:1646–1655.
20. Siegenthaler KO, Künkel A, Skupin G, Yamamoto M (2011)
with a chain extender. Pristine PBAT, which has high rate Adv Polym Sci:1–46
of degradation in soil, has suffered strong reduction on this 21. Ma P, Cai X, Zhang Y, Wang S, Dong W, Chen M, Lemstra PJ
rate by the presence of the chain extender, which reflected (2014) Polym Degrad Stab 102:145–151
in the behavior of the blends. The chain extender func- 22. Signori F, Boggioni A, Righetti MC, Rondán CE, Bronco S,
Ciardelli F (2015) Macromol Mater Eng 300:153–160
tional groups reacted with the groups generated during the 23. Dong W, Zou B, Yan Y, Ma P, Chen M (2013) Int J Mol Sci
polymers degradation in a competitive effect with molecu- 14:20189–20203
lar weight reduction, causing biodegradation delay. The 24. Schneider J, Manjure S, Narayan R (2016) J Appl Polym Sci
blends showed intermediate behavior compared to original 43310:1–9
25. Villalobos M, Awojulu A, Greeley T, Turco G, Deeter G (2006)
polymers, and the polymeric matrix behavior predominated Energy 31:3227–3234
until the surface of the dispersed phase was available for 26. Scheirs J, Long TE, Modern Polyesters: Chemistry and Technol-
the biodegradation process. Other factors, such as increased ogy of Polyesters and Copolyesters (Wiley Series in Polymer
crystallinity in the blends and morphology had also influ- Science. Australia, 2005)
27. Ghanbari A, Heuzey MC, Carreau PJ, Ton-That M (2013) Poly-
enced on the biodegradation process. The expectation of mer 54:1361–1369
accelerating PLA biodegradation in soil by compatibiliza- 28. Duarte S, Tavares AA, Lima PS, Andrade DL, Carvalho LH,
ted blending with PBAT was not achieved. Canedo EL, Silva SM (2016) Polym Degrad Stab 124:26–34.

13
J Polym Environ (2018) 26:330–341 341

29. Agarwal M, Koelling KW, Chalmers JJ (1998) Biotechnol Progr of soil analysis -Part 2—Chemical and microbiological proper-
14:517–526. ties (1982).
30. Kale G, Auras R, Singh SP (2007) Packag Technol Sci 20:49–70. 45. American Society for Testing and Materials (2012) ASTM

31. Tokiwa Y, B. P (2006) Calabia. Appl Microbiol Biotechnol
Standard D5988-12. Philadelphia, PA
72:244–251 46. Saadi Z, Cesar G, Bewa H, Benguigui L (2013) J Polym Environ
32. Oyama HT, Tanaka Y, Hirai S, Shida S, Kadosaka A (2011) J 21:893–901.
Polym Sci Part B 49:342–354. 47. Ho KG, Pometto AL (1999) J Polym Environ 7 101–108.
33. Nieddu E, Mazzucco L, Gentile P, Benko T, Balbo V, Mandrile 48. Tham WL, Poh BT, Mohd Ishak ZA, Chow WS (2015) J Polym
R, Ciardelli G (2009) React Funct Polym 69:371–379 Environ 23:242–250
34. Souza PMS, Corroque NA, Morales AR, Marin-Morales M, Mei 49. Auras R, Harte B, Selke S (2004) Macromol Biosci 4:835–864
LHI (2013) J Polym Environ 21:1052–1063 50. Pan P, Zhu B, Inoue Y (2007) Macromolecules 40:9664–9671
35. Souza PMS, Morales AR, Marin-Morales M, Mei LHI (2013) J 51. Di Lorenzo ML, Cocca M, Malinconico M (2011) Thermochim
Polym Environ 21:738–759 Acta 522:110–117.
36. Souza PMS, Morales AR, Mei LHI (2014) Polímeros
52. Tábi T, Sajó IE, Szabó F, Luyt AS, Kovács JG (2010) Polym Lett
24:110–116. 4:659–668
37. Kumar S, Maiti P (2015) Polymer 76:25–33 53. Carrasco F, Pagès P, Gámez-Pérez J, Santana OO, Maspoch ML
38. Stloukal P, Kalendova A, Mattausch H, Laske S, Holzer C,
(2010) Polym Degrad Stab 95:116–125
Koutny M (2015) Polym. Test 41:124–132 54. Kijchavengkul T, Auras R, Rubino M, Ngouajio M, Fernandez
39. Al-Itry R, Lamnawar K, Maazouz A (2014) Rheol Acta
RT (2008) Chemosphere 71:942–953
53:501–517 55. Henton DE, Gruber P, Lunt J, Randall J (2005) In: Mohanty
40. Arruda LC, Magaton M, R. E (2015) S. Bretas; M. M Ueki AK, Misra M, Drzal LT (eds) Taylor & Francis, Boca Raton,
Polym Test 43:27–37 pp. 527–577
41. Tabasi RY, Ajji A (2015) Polym Degrad Stab 120:435–442. 56. Lunt J (1998) Polym Degrad Stab 59:145–152
42. Weng YX, Jin YJ, Meng QY, Wang L, Zhang M, Wang Polym 57. Fukushima K, Tabuani D, Camino G (2009) Mater Sci Eng C
Y-Z (2013) Test 32:918–926 29:1433–1441
43. Kuchnier CN Study of the effect of multifunctional chain
58. Yasuniwa M, Tsubakihara S, Iura K, Ono Y, Dan Y, Takahashi K
extender in PLA/PBAT blends (State University of Campinas, (2006) Polymer 47:7554
Campinas, 2014).
44. Anderson JPE, In: Soil Respiration, (eds.) AL Page, RH Miller,
DR Keeney (Madison, Wisconsin, 1982), pp. 831–866. Methods

13

Potrebbero piacerti anche