You are on page 1of 6

RSC Advances

View Article Online


PAPER View Journal | View Issue

From waste to functional additives: thermal


stabilization and toughening of PVA with lignin
Published on 26 January 2016. Downloaded by Sichuan University on 9/17/2020 8:05:12 AM.

Cite this: RSC Adv., 2016, 6, 13797


Xiao-Qin Hu,ab De-Zhan Ye,c Jin-Bo Tang,ab Lin-Jie Zhangab and Xi Zhang*ab

Blend films of poly(vinyl alcohol) (PVA) and a graft copolymer (GL) of acrylic acid (AA) with eucalypt lignosulfonate
calcium (HLS) were prepared by using a solution casting method. The structure of GL/PVA blend films was
confirmed by X-ray diffraction (XRD), Fourier-transform infrared (FTIR) spectroscopy and scanning electron
microscopy (SEM), which indicate that GL/PVA is a homogeneous system due to the strong interactions
between PVA and GL. Differential scanning calorimetry (DSC) results indicate that only one glass transition
temperature (Tg) can be seen over the entire blending ratio. Meanwhile, the Tg was enhanced and the melting
point (Tm) was depressed when GL was added to PVA. With the addition of 3 wt% GL, the onset
Received 10th December 2015
Accepted 20th January 2016
decomposition temperature (To) was increased by 102  C compared to that of PVA. Compared with pure
PVA, the GL/PVA exhibited a remarkable improvement in mechanical properties: the tensile strength and
DOI: 10.1039/c5ra26385a
Young’s modulus of GL/PVA with 5 wt% GL were 39% and 285% higher than that of pure PVA, respectively.
www.rsc.org/advances These results show that a new melt-processing method of PVA may be developed by the addition of GL in PVA.

Usually, the maximum (Tmax) or onset decomposition (Tonset)


1 Introduction temperatures are improved by no more than 40  C, meanwhile
Poly(vinyl alcohol) (PVA) is widely used in many applications such remarkable deterioration of mechanical properties is commonly
as medical treatments, construction materials, and household observed. Interestingly, recently enhancement of the thermal
industries because of its low cost and many excellent properties properties and toughness of PVA was synchronously achieved
including mechanical performance, solvent resistance, biocom- with a low loading of melamine via hydrogen bond self-
patibility and biodegradability.1 Until now, two technologies, assembly.12 It was demonstrated that Tmax and Tonset of PVA were
thermal processing and solution processing, have been adopted increased by 13  C and 16  C, respectively, with only 0.5 wt%
for PVA. Despite its excellent water solubility, the casting proce- melamine addition; due to the formation of a physically cross-
dure of PVA is still time-consuming and uneconomical. Thus, linked network via H-bonding, the tensile strength and tough-
there is particular interest in obtaining PVA products through ness of PVA were improved by 22% and 200%, respectively.
melt extrusion or injection. Because of the strong inter- or Lignin, the second most abundant biomacromolecule that
intramolecular hydrogen bonding in PVA, its thermal processing exists in the plant kingdom, is relatively inexpensive and widely
window is very narrow, resulting in simultaneous decomposition available.13 Nevertheless, most lignin is burned as a low cost fuel,
during the extrusion process.2 Lowering PVA melting point by the leading to the waste of resources and growing environmental
incorporation of plasticizers could resolve this problem to some problems.14 Accordingly, great endeavors have been made to
extent.3,4 Nevertheless, this approach is unsatisfactory for entirely research its broader applications. Yeo et al. used modied lignin
overcoming the problem of the thermal decomposition of PVA as a ller in polypropylene composites.15 Binary blends of alka-
without sacricing it’s physical properties. In contrast to plasti- line lignin with three poly(vinyl alcohol) samples were investi-
cization, the improvement of PVA thermal stability with llers gated.16 Eucalyptus lignosulfonate calcium (HLS)/PVA blends
might offer a promising way for achieving its melt extrusion. For were researched by Ye, but the compatibility between them was
example, magnesium chloride, calcium oxide and silica show limited.17 In this work, a gra copolymer of acrylic acid (AA) and
a thermal retardation effect on PVA.2,5,6 Similar results were also eucalypt lignosulfonate calcium with abundant aromatic struc-
observed with the addition of nano-llers.7–11 However, their tures and hydroxyl groups was prepared14 and used to improve
effects on improving PVA thermal stability are rather limited. the thermal stability of PVA via H-bonding interactions.

a
Polymer Research Institute, Sichuan University, Chengdu 610065, PR China. E-mail:
zhangxi6352@163.com; Fax: +86-028-85402465; Tel: +86-028-85402465 2 Experimental
b
State Key Laboratory of Polymer Material Engineering, Sichuan University, Chengdu 2.1 Materials
610065, PR China
c
School of Materials Science and Engineering, Wuhan Textile University, 430200, PR Poly(vinyl alcohol) (PVA : DP ¼ 1700, degree of hydrolysis 99%)
China was provided by Sichuan Vinylon Factory, SINOPEC (China).

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 13797–13802 | 13797
View Article Online

RSC Advances Paper

Eucalyptus lignosulfonate calcium (HLS, 96%) was purchased


from Aladdin (China). Acrylic acid (AA), potassium persulfate
(K2S2O8), absolute ethyl alcohol and hydroquinone were
provided by Kelong Chemicals Co. Ltd. (China).
The gra copolymer (GL) of AA with HLS was prepared using
the same methods as our previous paper14 and the experimental
conditions are shown in Table 1.

2.2 Film preparation


Blend lms were prepared by a solution casting method. A
Published on 26 January 2016. Downloaded by Sichuan University on 9/17/2020 8:05:12 AM.

calculated amount of PVA and GL were dissolved in distilled


water at 90  C for 3 h. The dissolved solution was casted onto
PTFE Petri dishes and dried at 70  C in a vacuum oven for 15 h
to completely eliminate water. The dried lms were named as
GL/PVA and labelled with the prexes 5, 15, 25, 35, 50, which
represent the weight ratio of lignin gra copolymer to the total
Fig. 1 The FT-IR spectra of PVA, 5GL/PVA, and 15GL/PVA.
sample dry weight. For example, 5GL/PVA represents 5 wt% GL
in the lm (weight is based on the total quantity of dry PVA and
lignin gra copolymer).
(TGA, model SDT Q600). A sample (3–4 mg) was equilibrated at
30  C and then was heated from 30 to 600  C at a heating rate of
2.3 Characterization
10  C min1. The ow rate of the nitrogen atmosphere was
Before characterization, except for mechanical tests, all the maintained at 100 ml min1 during the whole process.
samples were dried under vacuum (P < 0.3 MPa) at 105  C for 5 h 2.3.6 Mechanical measurements. The tensile strength and
to eliminate water. elongation at break of PVA and blend lms were tested at room
2.3.1 Fourier-transform infrared (FT-IR) spectroscopy. temperature using a tensile tester (Instron 5567). Crosshead
FTIR spectra of samples were measured using a FT-IR spectro- speed was set at 20 mm min1. The initial gauge length of the
photometer (Nicolet 560) using attenuated total reection. The specimen was 20 mm. The width of each tensile sample is 4
measurements were carried out at 4 cm1 resolution with 32 mm. Thicknesses of the lms were measured with a micrometer
scans in the frequency of range of 4000–650 cm1. in triplicate. The lms were stored at 25  C with RH of 54% for 1
2.3.2 Differential scanning calorimetry (DSC). Glass tran- week before testing. The data for each sample was calculated
sition temperature (Tg) and melting temperature (Tm) of the from the average value of ve specimens. Tensile toughness (Ut)
blends were determined by using a Mettler Toledo DSC1 (DSC, can be calculated by integrating the area under the tensile
Mettler Corp. Switzerland) with a scan rate of 10 K min1 over curves, as expressed by following equation.12
a temperature range from 40  C to 300  C. The ow rate of the X
n
nitrogen atmosphere was kept at 50 ml min1 during the Ut ¼ s3i
process. i¼0

2.3.3 X-ray diffraction measurements (XRD). X-ray diffrac- where s and 3 respectively refer to the tensile stress and strain at
tion patterns were recorded in an angular range of 3–50 (2q) by failure.
using an X’Pert Pro XRD diffractometer equipped with Cu-Ka
radiation operated at 50 kV and 35 mA. The measurements were
performed at a scanning speed of 2q ¼ 0.06 s1. 3 Results and discussion
2.3.4 Scanning electron microscope (SEM). The morphol-
ogies of fracture surfaces were analyzed by SEM (JSM-5900LV) in 3.1 FTIR analysis
order to identify the compatibility between GL and PVA in the FT-IR spectroscopy is a sensitive tool for monitoring changes in
blends. Prior to characterization, blends were immersed in the interactions of blends.18 The FTIR spectra of PVA and blends
liquid nitrogen for 3 h to obtain freeze-fractured surfaces. with the addition of lignin gra polymers are illustrated in
2.3.5 Thermogravimetric analysis (TGA). The onset (To) Fig. 1. For neat PVA, the wide absorption peak centered at 3313
and maximum decomposition (Tmax) temperatures of PVA and cm1 is related to the –OH stretching vibration due to the
blends were determined using thermal gravimetric analysis formation of intermolecular/intramolecular hydrogen bonding.

Table 1 Experimental conditions for grafting reactions of HLS and AA

Sample HLS (g) AA (g) H2O (ml) K2S2O8 (g) Hydroquinone (g) T ( C) Time (min)

GL 1.6 8 18.67 0.24 0.2 50 20

13798 | RSC Adv., 2016, 6, 13797–13802 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper RSC Advances


Published on 26 January 2016. Downloaded by Sichuan University on 9/17/2020 8:05:12 AM.

Fig. 2 DSC heating thermograms of PVA films as a function of GL content (heating rate ¼ 10 K min1).

polymer. In 5GL/PVA, this wide –OH absorption band is located


at 3305 cm1, declining by 8 cm1 in comparison with pure
PVA. Upon loading 15 wt% GL, it nally decreased to 3297 cm1,
which can be attributed to the interactions between the –OH
group and carbonyl group from lignin gra polymers. This
hypothesis could be proved by the changes in the absorption
peak of PVA emerging at 1736 cm1 which was assigned to the
typical peak of ester carbonyl (–C]O) structures. With
increasing loading of GL, this peak absorption showed similar
behavior that the –OH band did, shiing gradually from 1736
cm1 to 1728 cm1. This indicates that –C]O groups in PVA
also could form interactions with –OH or –COOH groups of
modied lignin.

3.2 Thermal properties and glass transition temperature


Fig. 2 shows the DSC heating thermograms of PVA lms
Fig. 3 X-ray diffraction patterns of PVA films with different concen- according to the concentration of GL. It can be seen that there
trations of GL. were two chain segment motion platforms located at around
80  C and 130  C for pure PVA. Obviously, the former platform
Aer blending with the gra polymer from lignin and acrylic represents the movement of PVA chains in the amorphous
acid, a red shi of –OH peak absorption was observed, sug- region and the latter platform is mainly caused by the crystal-
gesting strong interactions between PVA and the lignin gra line relaxations of PVA.19,20 The pure PVA displayed a Tg of
74.9  C and amorphous GL displayed a Tg of 108.59  C (Fig. 2A).

Fig. 4 SEM micrographs of GLPVA blend membranes.

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 13797–13802 | 13799
View Article Online

RSC Advances Paper


Published on 26 January 2016. Downloaded by Sichuan University on 9/17/2020 8:05:12 AM.

Fig. 5 Thermal analysis of PVA, GL and GLPVA blending films: (A) TGA curves and (B) DTG curves.

Table 2 Comparison of nano-fillers and GL effect on PVA maximum a compatible system, and there are strong interactions between
decomposition temperature (Tmax) GL and PVA. This result is in accordance with our previous
results.
Enhanced Tmax ( C,
Sample compared with pure PVA)
3.4 Scanning electron microscopy
PVA/graphite oxide 5 wt%31 105
PVA/nano montmorillonite 5 wt%8 40 In order to conrm the compatibility of blends, the fracture
PVA/graphite oxide 3 wt%9 36 surfaces of pure PVA as well as GL/PVA with different compo-
PVA/graphite 3.5 wt%10 20–30 sitions (15GL/PVA and 50GL/PVA) were characterized by SEM
3 wt% GL/PVA 109
(Fig. 4). Pure PVA shows a very smooth fracture surface. Inter-
estingly, it is noticed that blend lms also show smooth and
homogeneous surfaces even with the incorporation of 50 wt%
Moreover, it can be clearly seen that only one single composi- GL. GL is homogeneously dispersed in the PVA matrix. As re-
tional dependent glass transition temperature (Tg) can be seen ported for PVA blends with alkali lignin, starch and cellulose
over the entire blend ratio for GL/PVA blends. It can be inter- and so on, the aggregation of particles and deteriorated
preted that the compatibility between GL and PVA is very good compatibility are commonly observed due to a lack of strong
and the system is homogeneous and single phase which is in interactions.27–29 Our previous studies demonstrated that PVA/
accordance with the results of SEM.21,22 Aer GL was incorpo- lignosulfonate blends with strong interactions (the q value in
rated, Tg was gradually enhanced. For the composite lm with the Kwei equation is 62.4  10.0) still become partly immis-
35 wt% GL, the Tg was increased to 94.1  C with an increment of cible with a 50 wt% addition of lignosulfonate.17 Nevertheless,
19.2  C, compared with pure PVA. This can be ascribed to the the absence of particle aggregation is observed in 50GL/PVA
strong hydrogen bonding interactions between GL and PVA, blend, indicating that the modication of lignosulfonate by
which can restrict the free movement and arrangement of PVA graing is an effective method for improving blend compati-
chains.12 Besides, the melting point (Tm) gradually decreased as bility. It could be explained by the fact that the inter-hydrogen
GL content was increased (from 229  C for pure PVA to 214  C bonding between PVA and lignin is enhanced by adopting this
for 35GL/PVA), which was in accordance with the previous simple graing method. This good compatibility is obviously
literature (Fig. 2B).17 The depression of Tm indicates that there benecial for achieving higher thermal stability and mechan-
may be strong interactions in the blend system.23,24 Thus, it can ical properties of blends, as proved by later discussions. SEM
be considered to be GL in the blend system that contributes to micrographs of GLPVA blend membranes.
the formation of strong hydrogen bonding with PVA molecules.

3.5 Thermogravimetric analysis


3.3 XRD analysis TG and DTG curves of pure PVA, GL and GL/PVA blends were
Within the angular range of 5–40  C, there are two main crys- recorded in order to study the effect of lignin gra polymer on
talline maxima which may be indexed as 101 at 2q ¼ 19.88 , and the thermal degradation behavior of PVA (Fig. 5). Interestingly,
200 at 2q ¼ 22.76 for PVA (Fig. 3). The former crystalline peak, the onset decomposition temperature (To) and the maximum
which is stronger, is oen used for qualitative judgment of the decomposition (Tmax) of PVA were remarkably improved with
variation of crystallinity for PVA.25 With rising GL content, this the addition of GL. To of pure PVA and GL are 239  C and 217  C,
diffraction peak decreased in intensity indicating that the respectively. By only adding 3 wt% or 5 wt% GL, To and Tmax
ordered arrangement of PVA main chains in the crystalline were signicantly improved by about 102  C and 109  C,
region could be disturbed by the incorporation of GL, the respectively, compared with pure PVA. With a further increase
crystallization process of PVA can be suppressed and the crys- of GL, To and Tmax were found to slightly decrease. As reported
tallinity can decrease.26 It can be inferred that GL/PVA is in the literature, 5 wt% of nano-montmorillonite improved the

13800 | RSC Adv., 2016, 6, 13797–13802 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper RSC Advances

melting peak moves to a lower temperature, from the TGA


results together with DSC proles shown in Fig. 2. The
temperature gap between To and Tm was increased, indicating
a new way of thermal processing of PVA. As we all know,
hydroxyl groups take part in the thermal degradation of PVA,
and the stability of hydroxyl groups is crucial for the stabiliza-
tion of PVA.30 A possible explanation for the high thermal
stability of GL/PVA blends is the introduction of aromatic
structures of lignin into PVA chains via strong hydrogen
bonding. Moreover, PVA blends even with high compositions of
Published on 26 January 2016. Downloaded by Sichuan University on 9/17/2020 8:05:12 AM.

GL still showed good compatibility, as proved by SEM and DSC


results. Notably, these functional llers are derived from bio-
based resources which are rather inexpensive and environ-
mentally friendly.

3.6 Mechanical properties


Fig. 6(A)–(C) show the mechanical properties of the blend lms.
Pure PVA displays a tensile strength of 49.90 MPa, Young’s
modulus of 0.71 GPa, elongation at break of 225.85% and
a tensile toughness of 89.86 MJ m3. However, low addition of
GL had signicant effect on the mechanical properties of PVA.
With the incorporation of 5 wt% GL, almost all mechanical
parameters achieved maximum values. The tensile strength and
toughness were increased to 69.15 MPa and 130.19 MJ m3,
respectively, 39% and 45% higher than those of pure PVA.
Meanwhile the elongation at break was slightly deceased to
218.86% and the Young’s modulus was sharply increased to
2.73 GPa. When the GL content reached 15 wt%, the Young’s
modulus achieved the maximum value (3.28 GPa), 362% higher
than that of pure PVA. With a further increase of GL, the
mechanical parameters gradually decreased. At the 35 wt% GL
level, the tensile strength and elongation at break decreased to
44.91 MPa and 125.42%, respectively. But the Young’s modulus
was 196% higher than that of pure PVA. Su et al. reported that
when the mass ratio of alkali and PVA was 2 : 10, the tensile
strength and elongation at break were increased by 20% and
11%, respectively, compared with PVA.27 Ye et al. recently
observed that lignosulfonate (HLS)/PVA blend lms also had
better mechanical properties than pure PVA lm due to the
formation of strong hydrogen bonding and the rigidity of lignin
structures.17 Moreover, adding a small amount of MA can
signicantly improve the strength, modulus, and toughness of
PVA due to the formation of a physically crosslinked network via
H-bond self-assembly.12 Hence, the observed enhancements can
Fig. 6 (A) Tensile stress–strain curves, (B) tensile strength and Young’s be primarily attributed to the strong hydrogen bonding between
modulus and (C) elongation at break and tensile toughness at break for GL and PVA and the rigidness of lignin particles.
PVA and its blends as a function of GL content.

4 Conclusion
Tmax of PVA by 40  C and introduction of 3 wt% graphene oxide
In this paper, modied lignin (GL) was used as a reinforcer for
also gave the same level of improvement in the Tmax of PVA(36
 PVA and it was amazing that the properties of PVA were
C), shown in Table 2.8,9 To the best of our knowledge, GL/PVA
considerably changed. GL was homogeneously dispersed in the
blends show a much better thermal stability compared with any
PVA matrix and resulted in the enhancement of many proper-
other results from PVA blends with inorganic or organic llers.
ties. The results of DSC show that only one Tg can be seen over
Aer GL is incorporated, it is obvious that the onset decompo-
the entire blend ratio indicating that the blend systems are
sition temperature moves to a higher temperature, while the
homogeneous and miscible. The obtained TGA and DTG results

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 13797–13802 | 13801
View Article Online

RSC Advances Paper

suggest a very small amount (3% wt) of GL signicantly alters 13 S. Kubo and J. F. Kadla, Biomacromolecules, 2003, 4, 561–567.
the thermal properties of PVA and the thermal decomposition 14 D. Z. Ye, L. Jiang, C. Ma, M.-H. Zhang and X. Zhang, Int. J.
temperature can be higher than that of pure PVA by about 102 Biol. Macromol., 2014, 63, 43–48.

C, which was attributed to strong hydrogen bonding. The 15 J.-S. Yeo, D.-W. Seong and S.-H. Hwang, J. Ind. Eng. Chem.,
mechanical properties of the blend lms were also obviously 2015, 31, 80–85.
improved. With addition of 5 wt% GL, the tensile strength and 16 E. Corradini, E. A. G. Pineda and A. A. W. Hechenleitner,
Young’s modulus were 39% and 285% higher than those of pure Polym. Degrad. Stab., 1999, 66, 199–208.
PVA. Moreover, the tensile toughness was also improved. To 17 D.-Z. Ye, L. Jiang, X.-Q. Hu, M.-H. Zhang and X. Zhang, Int. J.
summarize, these results can be attributed to the strong Biol. Macromol., 2016, 83, 209–215.
hydrogen bonding between GL and PVA. This work may open 18 E. Sheha, M. Nasr and M. El-Mansy, Phys. Scr., 2013, 88,
Published on 26 January 2016. Downloaded by Sichuan University on 9/17/2020 8:05:12 AM.

a door to realize the melt processing of PVA. 035701.


19 R. M. Hodge, T. J. Bastow, G. H. Edward, G. P. Simon and
References A. J. Hill, Macromolecules, 1996, 29, 8137–8143.
20 Y. Nishio and R. S. J. Manley, Macromolecules, 1988, 21,
1 P. Liu, W. Chen, Y. Liu, S. Bai and Q. Wang, Polym. Degrad. 1270–1277.
Stab., 2014, 109, 261–269. 21 X. Lu and R. A. Weiss, Macromolecules, 1992, 25, 3242–3246.
2 X. Jiang, T. Jiang, X. Zhang, H. Dai and X. Zhang, Polym. Eng. 22 P. C. Painter, J. F. Graf and M. M. Coleman, Macromolecules,
Sci., 2012, 52, 2245–2252. 1991, 24, 5630–5638.
3 J. Jang and D. K. Lee, Polymer, 2003, 44, 8139–8146. 23 C. Sawatari and T. Kondo, Macromolecules, 1999, 32, 1949–
4 W. Wu, H. Tian and A. Xiang, J. Polym. Environ., 2012, 20, 63– 1955.
69. 24 Y. Nishio, T. Haratani, T. Takahashi and R. S. J. Manley,
5 A. R. Butt, S. Ejaz, J. C. Baron, M. Ikram and S. Ali, Digest Macromolecules, 1989, 22, 2547–2549.
Journal of Nanomaterials and Biostructures, 2015, 10, 799–809. 25 R. F. Bhajantri, V. Ravindrachary, A. Harisha, V. Crasta,
6 J. Dodda, P. Bělský, J. Chmelař, T. Remiš, K. Smolná, S. P. Nayak and B. Poojary, Polymer, 2006, 47, 3591–3598.
M. Tomáš, L. Kullová and J. Kadlec, J. Mater. Sci., 2015, 50, 26 C.-H. Chen, F.-Y. Wang, C.-F. Mao, W.-T. Liao and
6477–6490. C.-D. Hsieh, Int. J. Biol. Macromol., 2008, 43, 37–42.
7 W. Zhang, X. He, C. Li, X. Zhang, C. Lu, X. Zhang and 27 L. Su, Z. Xing, D. Wang, G. Xu, S. Ren and G. Fang,
Y. Deng, Cellulose, 2014, 21, 485–494. BioResources, 2013, 8(3), 3532–3543.
8 Y.-H. Yu, C.-Y. Lin, J.-M. Yeh and W.-H. Lin, Polymer, 2003, 28 X. Jiang, H. Li, Y. Luo, Y. Zhao and L. Hou, Int. J. Biol.
44, 3553–3560. Macromol., 2016, 82, 223–230.
9 Y. Xu, W. Hong, H. Bai, C. Li and G. Shi, Carbon, 2009, 47, 29 N. Hameed, R. Xiong, N. Salim and Q. Guo, Cellulose, 2013,
3538–3543. 20, 2517–2527.
10 X. Yang, L. Li, S. Shang and X.-M. Tao, Polymer, 2010, 51, 30 X. Jiang, B. Tan, X. Zhang, D. Ye, H. Dai and X. Zhang, J. Appl.
3431–3435. Polym. Sci., 2012, 125, 697–703.
11 S. G. Chaudhri, B. H. Rajai and P. S. Singh, RSC Adv., 2015, 5, 31 H. J. Salavagione, G. Martinez and M. A. Gomez, J. Mater.
65862–65869. Chem., 2009, 19, 5027–5032.
12 P. A. Song, Z. Xu and Q. Guo, ACS Macro Lett., 2013, 2, 1100–
1104.

13802 | RSC Adv., 2016, 6, 13797–13802 This journal is © The Royal Society of Chemistry 2016

You might also like