Next Article in Journal
An Easy Route to Wettability Changes of Polyethylene Terephthalate–Silicon Oxide Substrate Films for High Barrier Applications, Surface-Modified with a Self-Assembled Monolayer of Fluoroalkylsilanes
Next Article in Special Issue
Water Softening Using a Light-Responsive, Spiropyran-Modified Nanofiltration Membrane
Previous Article in Journal
Hydrophobicity of Polyacrylate Emulsion Film Enhanced by Introduction of Nano-SiO2 and Fluorine
Previous Article in Special Issue
Synthesis and Characterization of BaTiO3/Polypyrrole Composites with Exceptional Dielectric Behaviour
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Green Polyurethanes from Renewable Isocyanates and Biobased White Dextrins

1
Macromolecular Chemistry and New Polymeric Materials, Zernike Institute for Advanced Materials, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
2
Dutch Polymer Institute (DPI), P.O. Box 902, 5600 AX Eindhoven, The Netherlands
*
Author to whom correspondence should be addressed.
Polymers 2019, 11(2), 256; https://doi.org/10.3390/polym11020256
Submission received: 11 September 2018 / Revised: 29 January 2019 / Accepted: 31 January 2019 / Published: 3 February 2019
(This article belongs to the Special Issue Polymers: Design, Function and Application)

Abstract

:
Polyurethanes (PUs) are an important class of polymers due to their low density and thermal conductivity combined with their interesting mechanical properties—they are extensively used as thermal and sound insulators, as well as structural and comfort materials. Despite the broad range of applications, the production of PUs is still highly petroleum-dependent. The use of carbohydrates in PU synthesis has not yet been studied extensively, even though, as multihydroxyl compounds, they can easily serve as crosslinkers in PU synthesis. Partially or potentially biobased di-, tri- or poly-isocyanates can further be used to increase the renewable content of PUs. In our research, PU films could be easily produced using two bio-based isocyanates—ethyl ester L-lysine diisocyanate (LLDI] and ethyl ester l-lysine triisocyanate (LLTI)—, one commercial isocyanate—isophorone diisocyanate (IPDI), and a bio-based white dextrin (AVEDEX W80) as a crosslinker. The thermal and mechanical properties are evaluated and compared as well as the stability against solvents.

Graphical Abstract

1. Introduction

Polyurethanes (PUs) are versatile polymers, created from the reaction between a polyol and a polyisocyanate, that are used in a large range of applications, such as automotive, furniture, construction, footwear, etc. The PU industry is strongly dependent on fossil-based polyols and polyisocyanates. Developing novel sustainable polyols from valuable biobased building blocks is a first step toward strong and durable development [1,2,3].
In recent years, the preparation of PUs from renewable sources such as vegetable oil-based materials [4], lignin [5], limonene [6], and even coffee grounds [7] has been receiving increasing attention because of economic and environmental concerns [8,9,10,11,12,13]. The use of carbohydrates in PU synthesis has not yet been studied extensively, even though, as multihydroxyl compounds [14,15,16,17,18,19], they can easily serve as crosslinkers in PU synthesis. Furthermore, they can impart mechanical strength, biodegradability, and biocompatibility to the produced PUs. Carbohydrates are reported to be embedded in PU networks by using them as composites/fillers and by covalent linkage with isocyanate to form crosslinked networks [20]. Solanki et al., for instance, synthesized castor oil-based PUs crosslinked with starch and reported excellent mechanical properties of the produced materials [21]. Carbohydrates, such as cellulose [22,23,24,25,26] or starch [27] nanocrystals, have been used as useful fillers in PUs. We have recently reported PU synthesis and film formation capacity using an asymmetric cyclic aliphatic diisocyanate and acetylated and pristine partially hydrolyzed amylopectin/white dextrin (AVEDEX W80) as a crosslinker [28,29].
The next logical step towards a greener PU formulation is then, of course, the use of partially or potentially biobased di-, tri- or poly-isocyanates. While synthetic pathways to biobased isocyanates still require phosgene as a petroleum-based reagent, several commercial isocyanates with high renewable content exist, such as isocyanates based on fatty acids or amino acids.
l-lysine-based diisocyanates have frequently been used for biomedical PUs in the last several years as drug delivery systems, hydrogels, implant materials, etc. [30,31,32,33,34,35,36]. In some very inspiring recent research, the L-lysine based ethyl ester l-lysine diisocyanate was compared to petrol-based hexamethylene diisocyanate, and isophorone diisocyanate in terms of reactivity and final properties of the PU [37,38,39,40,41] and even fully renewable thermoplastic poly(ester urethane urea)s were produced [42]. More studies on enzymatic breakdown/biodegradability of (biobased) PUs are conducted in recent years proving the importance of environmentally friendly PU applications [9,43,44,45,46,47].
In our previous work, we found that pristine AVEDEX W80 resulted in the best PU films [29,48] (i.e., with no cracks, inhomogeneity, or swelling) and we, therefore, continue with this system and two bio-based isocyanates—ethyl ester l-lysine diisocyanate (LLDI) and ethyl ester l-lysine triisocyanate (LLTI)—and one commercial isocyanate—isophorone diisocyanate (IPDI). The optimal reaction conditions in terms of casting time and temperature, isocyanate used, amount of solvent, and excess of isocyanate were also determined.

2. Materials and Methods

Already degraded amylopectin–white dextrin (AVEDEX W80) was obtained from AVEBE (Veendam, The Netherlands) and used without further purification. Iodine was provided by Boom and purified by sublimation twice. Isophorone diisocyanate (IPDI) was provided by Merck (Darmstadt, Germany) and ethyl ester L-lysine diisocyanate (LLDI) and ethyl ester l-lysine triisocyanate (LLTI) were provided by Shanghai Infine Chemicals Co. Ltd. (Shanghai, China). Acetic anhydride ((CH3CO)2O) (≥99.0%) and propionic anhydride ((CH3CH2CO)2O) (≥96.0%) were provided by Fluka (Bucharest, Romania), sodium thiosulfate (NaS2O3) (≥98.0%) and DMSO (99%) were provided by Sigma-Aldrich (St. Louis, MO, USA) and ethanol (≥99%) was purchased from Merck (Darmstadt, Germany) and used without further purification.
Attenuated total reflection-Fourier transform infrared (ATR-FTIR) measurements were characterized by a Bruker IFS88 FT-IR spectrometer (Leiderdorp, The Netherlands). For each sample, 128 scans were performed.
Thermal transitions were measured by differential scanning calorimetry (DSC) on a TA-Instruments Q1000 (Hertfordshire, UK). The heating and cooling rates were 10 °C/min. To remove the remaining water and solvents in the polymer, the tested sample was first heated to 100 °C at 10 °C/min, maintained at this temperature for 5 min, and then cooled to room temperature before the standard DSC measurement. Each sample was run for 4 cycles, and the Tg was calculated from the second.
Specimens were made from PU samples and were tested on a tensile tester to study their mechanical properties, such as stress–strain behavior and Young’s modulus. The samples were kept between two Teflon plates to ensure a smooth and straight sample form, especially samples that had the tendency to break easily. Mechanical properties were tested on an Instron Tensile Tester (Boechout, Belgium) with a pulling speed of 10 mm/min. The distance between the clamps was 25 mm with a pressure of 7 bar.

2.1. Crosslinking of Acylated Polyols with Di- and Tri-Iisocyanates

Point two gram polyol AVEDEX W80 was dissolved in 7.5 mL of solvent (DMSO). A 1.05-fold excess of a di- and/or tri-isocyanate (i.e., a 1.05-fold excess of –NCO groups to –OH groups) was added, the mixture was stirred for 1 to 2 min, and the resulting homogenous mixture was poured onto a petri dish. The petri dish was placed on a heating block at a defined temperature (120 °C), covered with filter paper and the heating block was sealed to provide a solvent atmosphere. After a defined period of time, the petri dish was removed from the heating block.

2.2. Resistance against Organic Solvents

To investigate the resistance of the obtained polyurethane films against organic solvents, samples were weighed, after drying to constant weight, before and after immersing them in an organic solvent.

3. Results

The synthesis of polyurethane films by the crosslinking of carbohydrates as polyol components with isocyanates—see Scheme 1—was successfully conducted. The selection of isocyanates available for PU synthesis with polysaccharide polyols is almost unrestricted. As aromatic isocyanates might introduce a color into the final material, we did not consider them here. Two bio-based isocyanates—ethyl ester l-lysine diisocyanate (LLDI) and ethyl ester l-lysine triisocyanate (LLTI)—and one commercial isocyanate—isophorone diisocyanate (IPDI)—were used.
PUs derived from the diisocyanates IPDI and LLDI were too hard and brittle to test after drying. LLTI-based PUs behaved more like a rubber than a plastic; after drying the material became hard, but it was still flexible. PUs from IPDI always resulted in a hard and brittle material (after drying), but they are the only ones that are transparent due to the yellow color of the biobased isocyanates. PUs from LLDI also became brittle after drying, but the material was soft and easy to scratch.
FT-IR spectra of all films show successful PU synthesis; Figure 1 shows the FT-IR spectrum of an IPDI-based PU. The peaks of the amide I and amide II bands from the newly formed polyurethane linkages are visible at 1550 cm−1 and 1650 cm−1 respectively. All materials are not fully reacted, i.e., not all of the OH-functions of the polyol formed urethane linkages with the di- and tri-isocyanates, not even while using a 50% excess of isocyanate. It seems to be impossible to convert all free OH groups in pristine AVEDEX W80 as crosslinking becomes too high. In our previous work, we also observed that PU films with an excessively high crosslinking become brittle so that a full conversion is not desirable from the start.
To improve the properties of synthesized PUs, we decided to perform experiments with mixtures of isocyanates from this series. Two series of experiments with different ratios of two isocyanates were carried out, with the diisocyanate always the main component of the isocyanate and up to 50% (by volume) LLTI as a “co”-isocyanate. Combinations of the two diisocyanates were also prepared, with different ratios of these two isocyanates.
Mechanical properties of the produced films were tested on an Instron Tensile Tester with a pulling speed of 10 mm/min and are summarized in Table 1—the stress–strain curves are shown in Figure 2, Figure 3 and Figure 4.
Gradual replacement of the hard diisocyanate with the soft triisocyanate resulted in a softer, less rigid and breakable material. At 29% LLTI, the strain at maximum load was twice as large as at 12% LLTI and samples were much more flexible. The curve progression was straight at 8% LLTI, but at 29% LLTI the curve already looked much more like a stress–strain curve from a rubber, with a plateau at its maximum and much higher strain than at 8% LLTI (see Figure 2).
Shifting the ratio of two diisocyanates slightly from a 1:1 ratio toward either side did not change the mechanical properties of the material. As seen in Table 1, neither the average modulus, nor the stress or the strain at max load change significantly. Only the curve progression becomes more rubbery, like in the sample with 58% LLDI.
Gradually replacing LLDI with LLTI during solvent casting resulted in a much harder material compared to PUs consisting only of LLDI. At approximately 30% LLTI, the material has the most desirable properties, as the surface is hard, but the whole sample is still flexible enough to bend, and the strain before breaking is over 100%. The LLDI/LLTI PU is by far the one with the greatest strain at maximum load (Table 1).
Increasing the percentage of LLTI changes the PU properties even further, with the flexibility almost lost and the stress the sample can withstand before breaking very low compared with the other combinations of isocyanates. The maximum strain is also greatly reduced with an increase of LLTI above 30%.
The thermal properties of the produced PUs were assessed by differential scanning calorimetry. The highest Tg was observed for the IPDI-based PU (Tg = 157 °C). To investigate the dependence of Tg on the ratio of isocyanates, samples with different ratios were tested; this data is presented in Figure 5. In the case of the IPDI-based PUs. the effect is clearly visible; lowering the IPDI ratio lowers the Tg. In the case of the LLDI/LLTI-based PUs, change of the LLDI/LLTI ratio does not result in significant change in Tg; increasing the percentage of LLDI results in Tg being slightly shifted towards lower temperatures. At this point, we cannot explain the one outlier at around 80% LLDI content.
To test the resistance of the PU films against organic solvents, different samples were chosen and immersed in DMSO for 24 h. DMSO was chosen because it showed the most aggressive behavior toward the films. The difference in weight before and after immersion after drying until constant weight was calculated; this data is presented in Table 2. As observed by FT-IR measurements, the samples are not changed chemically during the solvent resistance test, indicating that the weight loss mainly results from washing away of unreacted polyol from the PU sample (as indicated by the lower OH-peak and higher amide I and II peaks). Samples made with IPDI seem to be more resistant against attacks from DMSO. The maximum weight loss in samples with IPDI is 7%, as compared to 12.5% for LLTI and almost 30% for LLDI.

4. Conclusions

Unmodified AVEDEX W80 was able to form homogenous films with the tested di- and tri-isocyanates, after optimization of several reaction conditions. With commercially used IPDI, the resulting material is transparent, rigid, and tends to break quite easily, especially after drying in the vacuum oven.
PUs derived from LLDI are flexible, but rather milky than transparent and they experience a coarse surface. They are more flexible than the IPDI based PUs but also tend to crack when fully dry as proven by tensile tests. As the biobased LLDI and LLTI are colored, the resulting films have a slightly yellow color (more pronounced in the PU films of LLTI).
Addition of the trifunctional isocyanate LLTI to either IPDI- or LLDI-based formulations resulted in a dry material, which has a high scratch resistance but is still flexible without breaking. Ratios of the isocyanates play an important role in the final properties of the material.
In general, the IPDI-based PUs shows higher stability in DMSO than the biobased. Additionally, the biobased PUs tend to swell in organic solvents, which is not the case for IPDI-based PUs. Besides differences in brittleness, differences in Tg can be observed in all four PUs. Combining two of these isocyanates results in changes in Tg, especially while replacing IPDI with LLDI, resulting in a drop in Tg. Gradually replacing LLDI with LLTI does not result in significant Tg changes in the material.
The mechanical properties of these materials depend on the isocyanate used for the PU formulation. If IPDI is the major component in the isocyanate mixture, the mechanical behavior is that of a hard and brittle material. Addition of the trifunctional isocyanate results in a much more rubbery material, visible in the stress–strain diagram. The slope of the curve starts to decrease at some point, reaching a plateau, which then leads to “necking” in the tensile tests.

Author Contributions

J.K. designed and performed the experiments reported in this publication and wrote parts of the paper. K.L. provided guidance, idea and suggestion to the research, helped in analyzing some data, revising and writing this paper.

Funding

This research forms part of the research programme of the Dutch Polymer Institute (DPI), project #673.

Acknowledgments

This research forms part of the research programme of the Dutch Polymer Institute (DPI), project #673.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

PUpolyurethane
AVEDEX W80amylopectin/white dextrin
LLDIethyl ester l-lysine diisocyanate
LLTIethyl ester l-lysine triisocyanate
IDPIisophorone diisocyanate

References

  1. Dworakowska, S.; Bogdal, D.; Zaccheria, F.; Ravasio, N. The role of catalysis in the synthesis of polyurethane foams based on renewable raw materials. Catal. Today 2014, 223, 148–156. [Google Scholar] [CrossRef]
  2. Kaiser, T.; Rathgeb, A.; Gertig, C.; Bardow, A.; Leonhard, K.; Jupke, A. Carbon2polymer-conceptual design of a co2-based process for the production of isocyanates. Chemie Ingenieur Technik 2018, 90, 1497–1503. [Google Scholar] [CrossRef]
  3. Wegener, G.; Brandt, M.; Duda, L.; Hofmann, J.; Klesczewski, B.; Koch, D.; Kumpf, R.J.; Orzesek, H.; Pirkl, H.G.; Six, C.; et al. Trends in industrial catalysis in the polyurethane industry. Appl. Catal. A Gen. 2001, 221, 303–335. [Google Scholar] [CrossRef]
  4. Chuayjuljit, S.; Maungchareon, A.; Saravari, O. Preparation and properties of palm oil-based rigid polyurethane nanocomposite foams. J. Reinf. Plast. Compos. 2010, 29, 218–225. [Google Scholar] [CrossRef]
  5. Bernardini, J.; Cinelli, P.; Anguillesi, I.; Coltelli, M.B.; Lazzeri, A. Flexible polyurethane foams green production employing lignin or oxypropylated lignin. Eur. Polym. J. 2015, 64, 147–156. [Google Scholar] [CrossRef]
  6. Gupta, R.K.; Ionescu, M.; Radojcic, D.; Wan, X.; Petrovic, Z.S. Novel renewable polyols based on limonene for rigid polyurethane foams. J. Polym. Environ. 2014, 22, 304–309. [Google Scholar] [CrossRef]
  7. Gama, N.V.; Soares, B.; Freire, C.S.R.; Silva, R.; Neto, C.P.; Barros-Timmons, A.; Ferreira, A. Bio-based polyurethane foams toward applications beyond thermal insulation. Mater. Des. 2015, 76, 77–85. [Google Scholar] [CrossRef]
  8. Noreen, A.; Zia, K.M.; Zuber, M.; Tabasum, S.; Zahoor, A.F. Bio-based polyurethane: An efficient and environment friendly coating systems: A review. Prog. Org. Coat. 2016, 91, 25–32. [Google Scholar] [CrossRef]
  9. Suthapakti, K.; Molloy, R.; Leejarkpai, T. Disintegration testing of biodegradable poly(l-lactide)/thermoplastic polyurethane melt blended films. Chiang Mai J. Sci. 2018, 45, 2079–2091. [Google Scholar]
  10. Uscategui, Y.L.; Diaz, L.E.; Valero, M.F. Biomedical applications of polyurethanes. Quimica Nova 2018, 41, 434–445. [Google Scholar] [CrossRef]
  11. Das, S.; Pandey, P.; Mohanty, S.; Nayak, S.K. Evaluation of biodegradability of green polyurethane/nanosilica composite synthesized from transesterified castor oil and palm oil based isocyanate. Int. Biodeterior. Biodegrad. 2017, 117, 278–288. [Google Scholar] [CrossRef]
  12. Nakhoda, H.M.; Dahman, Y. Mechanical properties and biodegradability of porous polyurethanes reinforced with green nanofibers for applications in tissue engineering. Polym. Bull. 2016, 73, 2039–2055. [Google Scholar] [CrossRef]
  13. Duguay, D.G.; Labow, R.S.; Santerre, J.P.; McLean, D.D. Development of a mathematical-model describing the enzymatic degradation of biomedical polyurethanes. 1. Background, rationale and model formulation. Polym. Degrad. Stab. 1995, 47, 229–249. [Google Scholar] [CrossRef]
  14. van der Vlist, J.; Schonen, I.; Loos, K. Utilization of glycosyltransferases for the synthesis of a densely packed hyperbranched polysaccharide brush coating as artificial glycocalyx. Biomacromolecules 2011, 12, 3728–3732. [Google Scholar] [CrossRef] [PubMed]
  15. Ciric, J.; Woortman, A.J.; Loos, K. Analysis of isoamylase debranched starches with size exclusion chromatography utilizing pfg columns. Carbohydr. Polym. 2014, 112, 458–461. [Google Scholar] [CrossRef] [PubMed]
  16. Ciric, J.; Rolland-Sabate, A.; Guilois, S.; Loos, K. Characterization of enzymatically synthesized amylopectin analogs via asymmetrical flow field flow fractionation. Polymer 2014, 55, 6271–6277. [Google Scholar] [CrossRef]
  17. Loos, K.; von Braunmuhl, V.; Stadler, R.; Landfester, K.; Spiess, H.W. Saccharide modified silica particles by enzymatic grafting. Macromol. Rapid Commun. 1997, 18, 927–938. [Google Scholar] [CrossRef]
  18. van der Vlist, J.; Palomo Reixach, M.; van der Maarel, M.; Dijkhuizen, L.; Schouten, A.J.; Loos, K. Synthesis of branched polyglucans by the tandem action of potato phosphorylase and deinococcus geothermalis glycogen branching enzyme. Macromol. Rapid Commun. 2008, 29, 1293–1297. [Google Scholar] [CrossRef]
  19. Mazzocchetti, L.; Tsoufis, T.; Rudolf, P.; Loos, K. Enzymatic synthesis of amylose brushes revisited: Details from x-ray photoelectron spectroscopy and spectroscopic ellipsometry. Macromol. Biosci. 2014, 14, 186–194. [Google Scholar] [CrossRef]
  20. Solanki, A.; Das, M.; Thakore, S. A review on carbohydrate embedded polyurethanes: An emerging area in the scope of biomedical applications. Carbohydr. Polym. 2018, 181, 1003–1016. [Google Scholar] [CrossRef]
  21. Solanki, A.; Mehta, J.; Thakore, S. Structure-property relationships and biocompatibility of carbohydrate crosslinked polyurethanes. Carbohydr. Polym. 2014, 110, 338–344. [Google Scholar] [CrossRef] [PubMed]
  22. Septevani, A.A.; Evans, D.A.C.; Annamalai, P.K.; Martin, D.J. The use of cellulose nanocrystals to enhance the thermal insulation properties and sustainability of rigid polyurethane foam. Ind. Crop. Prod. 2017, 107, 114–121. [Google Scholar] [CrossRef]
  23. Aranguren, M.I.; Marcovich, N.E.; Salgueiro, W.; Somoza, A. Effect of the nano-cellulose content on the properties of reinforced polyurethanes. A study using mechanical tests and positron anihilation spectroscopy. Polym. Test. 2013, 32, 115–122. [Google Scholar] [CrossRef]
  24. Auad, M.L.; Contos, V.S.; Nutt, S.; Aranguren, M.I.; Marcovich, N.E. Characterization of nanocellulose-reinforced shape memory polyurethanes. Polym. Int. 2008, 57, 651–659. [Google Scholar] [CrossRef]
  25. Wu, G.M.; Chen, J.; Huo, S.P.; Liu, G.F.; Kong, Z.W. Thermoset nanocomposites from two-component waterborne polyurethanes and cellulose whiskers. Carbohydr. Polym. 2014, 105, 207–213. [Google Scholar] [CrossRef] [PubMed]
  26. Wu, Q.J.; Henriksson, M.; Liu, X.; Berglund, L.A. A high strength nanocomposite based on microcrystalline cellulose and polyurethane. Biomacromolecules 2007, 8, 3687–3692. [Google Scholar] [CrossRef] [PubMed]
  27. Zou, J.W.; Zhang, F.; Huang, J.; Chang, P.R.; Su, Z.M.; Yu, J.H. Effects of starch nanocrystals on structure and properties of waterborne polyurethane-based composites. Carbohydr. Polym. 2011, 85, 824–831. [Google Scholar] [CrossRef]
  28. Konieczny, J.; Loos, K. Facile esterification of degraded and non-degraded starch. Macromol. Chem. Phys. 2018. [Google Scholar] [CrossRef]
  29. Konieczny, J.; Loos, K. Bio-based polyurethane films using white dextrins. J. Appl. Polym. Sci. 2018. [Google Scholar] [CrossRef]
  30. Fu, H.G.; Gao, H.; Wu, G.L.; Wang, Y.O.; Fan, Y.G.; Ma, J.B. Preparation and tunable temperature sensitivity of biodegradable polyurethane nanoassemblies from diisocyanate and poly(ethylene glycol). Soft Matter 2011, 7, 3546–3552. [Google Scholar] [CrossRef]
  31. Han, J.A.; Cao, R.W.; Chen, B.; Ye, L.; Zhang, A.Y.; Zhang, J.A.; Feng, Z.G. Electrospinning and biocompatibility evaluation of biodegradable polyurethanes based on l-lysine diisocyanate and l-lysine chain extender. J. Biomed. Mater. Res. Part A 2011, 96A, 705–714. [Google Scholar] [CrossRef] [PubMed]
  32. Marcos-Fernandez, A.; Abraham, G.A.; Valentin, J.L.; San Roman, J. Synthesis and characterization of biodegradable non-toxic poly(ester-urethane-urea)s based on poly(epsilon-caprolactone) and amino acid derivatives. Polymer 2006, 47, 785–798. [Google Scholar] [CrossRef]
  33. Mathew, S.; Baudis, S.; Neffe, A.T.; Behl, M.; Wischke, C.; Lendlein, A. Effect of diisocyanate linkers on the degradation characteristics of copolyester urethanes as potential drug carrier matrices. European Journal of Pharm. Biopharm. 2015, 95, 18–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Storey, R.F.; Wiggins, J.S.; Puckett, A.D. Hydrolyzable poly(ester-urethane) networks from l-lysine diisocyanate and d,l-lactide epsilon-caprolactone homopolyester and copolyester triols. J. Polym. Sci. Part A Polym. Chem. 1994, 32, 2345–2363. [Google Scholar] [CrossRef]
  35. Wang, Z.G.; Yu, L.Q.; Ding, M.M.; Tan, H.; Li, J.H.; Fu, Q.A. Preparation and rapid degradation of nontoxic biodegradable polyurethanes based on poly(lactic acid)-poly(ethylene glycol)-poly(lactic acid) and l-lysine diisocyanate. Polym. Chem. 2011, 2, 601–607. [Google Scholar] [CrossRef]
  36. Whang, C.H.; Lee, H.K.; Kundu, S.; Murthy, S.N.; Jo, S. Pluronic-based dual-stimuli sensitive polymers capable of thermal gelation and ph-dependent degradation for in situ biomedical application. J. Appl. Polym. Sci. 2018, 135. [Google Scholar] [CrossRef] [PubMed]
  37. Gustini, L.; Lavilla, C.; Finzel, L.; Noordover, B.A.J.; Hendrix, M.; Koning, C.E. Sustainable coatings from bio-based, enzymatically synthesized polyesters with enhanced functionalities. Polym. Chem. 2016, 7, 6586–6597. [Google Scholar] [CrossRef]
  38. Gustini, L.; Noordover, B.A.J.; Gehrels, C.; Dietz, C.; Koning, C.E. Enzymatic synthesis and preliminary evaluation as coating of sorbitol-based, hydroxy-functional polyesters with controlled molecular weights. Eur. Polym. J. 2015, 67, 459–475. [Google Scholar] [CrossRef] [Green Version]
  39. Li, Y.Y.; Noordover, B.A.J.; van Benthem, R.; Koning, C.E. Property profile of poly(urethane urea) dispersions containing dimer fatty acid-, sugar- and amine acid-based building blocks. Eur. Polym. J. 2014, 59, 8–18. [Google Scholar] [CrossRef]
  40. Li, Y.Y.; Noordover, B.A.J.; van Benthem, R.; Koning, C.E. Reactivity and regio-selectivity of renewable building blocks for the synthesis of water-dispersible polyurethane prepolymers. ACS Sustain. Chem. Eng. 2014, 2, 788–797. [Google Scholar] [CrossRef]
  41. Li, Y.Y.; Noordover, B.A.J.; van Benthem, R.; Koning, C.E. Bio-based poly(urethane urea) dispersions with low internal stabilizing agent contents and tunable thermal properties. Prog. Org. Coat. 2015, 86, 134–142. [Google Scholar] [CrossRef]
  42. Tang, D.L.; Thiyagarajan, S.; Noordover, B.A.J.; Koning, C.E.; van Es, D.S.; van Haveren, J. Fully renewable thermoplastic poly(ester urethane urea)s from bio-based diisocyanates. J. Renew. Mater. 2013, 1, 222–229. [Google Scholar] [CrossRef]
  43. Acik, G.; Kamaci, M.; Altinkok, C.; Karabulut, H.R.F.; Tasdelen, M.A. Synthesis and properties of soybean oil-based biodegradable polyurethane films. Prog. Org. Coat. 2018, 123, 261–266. [Google Scholar] [CrossRef]
  44. Joo, Y.S.; Cha, J.R.; Gong, M.S. Biodegradable shape-memory polymers using polycaprolactone and isosorbide based polyurethane blends. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 91, 426–435. [Google Scholar] [CrossRef] [PubMed]
  45. Loredo, A.; Arguello, A.; Rodriguez-Herrera, R.; Gutierrez-Sanchez, G.; Escamilla, A.; Aguilar, C. Fungal biodegradation of rigid polyurethane. Quimica Nova 2017, 40, 885–889. [Google Scholar]
  46. Suthapakti, K.; Molloy, R.; Punyodom, W.; Nalampang, K.; Leejarkpai, T.; Topham, P.D.; Tighe, B.J. Biodegradable compatibilized poly(l-lactide)/thermoplastic polyurethane blends: Design, preparation and property testing. J. Polym. Environ. 2018, 26, 1818–1830. [Google Scholar] [CrossRef]
  47. Zheng, L.C.; Li, C.C.; Xiao, Y.N.; Zhang, B.; Wang, Z.D. Biodegradable anti-fouling materials. Prog. Chem. 2017, 29, 824–832. [Google Scholar]
  48. Konieczny, J.; Loos, K. Polyurethane coatings based on renewable white dextrins and isocyanate trimers. Macromol. Rapid Commun. 2019. [Google Scholar] [CrossRef]
Scheme 1. Reaction scheme of the crosslinking of carbohydrates as polyol components with isocyanates.
Scheme 1. Reaction scheme of the crosslinking of carbohydrates as polyol components with isocyanates.
Polymers 11 00256 sch001
Figure 1. Fourier transform infrared (FT-IR) spectrum of isophorone diisocyanate (IPDI) based polyurethanes (PU) (75% IPDI/25% ethyl ester l-lysine triisocyanate (LLTI)).
Figure 1. Fourier transform infrared (FT-IR) spectrum of isophorone diisocyanate (IPDI) based polyurethanes (PU) (75% IPDI/25% ethyl ester l-lysine triisocyanate (LLTI)).
Polymers 11 00256 g001
Figure 2. Stress–strain behavior of PUs with different IPDI/LLTI ratios.
Figure 2. Stress–strain behavior of PUs with different IPDI/LLTI ratios.
Polymers 11 00256 g002
Figure 3. Stress–strain behavior of PUs with different IPDI/ ethyl ester L-lysine diisocyanate (LLDI) ratios.
Figure 3. Stress–strain behavior of PUs with different IPDI/ ethyl ester L-lysine diisocyanate (LLDI) ratios.
Polymers 11 00256 g003
Figure 4. Stress–strain behavior of PUs with different LLDI/LLTI ratios.
Figure 4. Stress–strain behavior of PUs with different LLDI/LLTI ratios.
Polymers 11 00256 g004
Figure 5. Relationship between Tg and isocyanate ratio of IPDI-based (left) and LLDI-based (right) PU films.
Figure 5. Relationship between Tg and isocyanate ratio of IPDI-based (left) and LLDI-based (right) PU films.
Polymers 11 00256 g005
Table 1. Average Young’s modulus, maximum stress and maximum strain of PUs with different isocyanate ratios.
Table 1. Average Young’s modulus, maximum stress and maximum strain of PUs with different isocyanate ratios.
% LLTI% IPDIYoung’s Modulus (MPa)Stress at Max Load (MPa)% Strain at Max Load
892112692
12881536332
25751554443
29711427465
% LLDI% IPDI
42581600443
50501501222
58421585413
% LLDI% LLTI
346622260
5050195692
6634182123
Table 2. Weight loss of different ethyl ester polyurethanes (PU) samples after immersion in DMSO for 1d.
Table 2. Weight loss of different ethyl ester polyurethanes (PU) samples after immersion in DMSO for 1d.
isocyanateIPDILLDILLDILLTILLDI/LLTIIPDI/LLTILLDI/LLTIIPDI/LLTILLDI/LLTILLDI/LLTILLDI/LLTI
ratio10010010010050/5066/3366/3375/2571/2992/879/21
weight before (g)0.05820.07500.07130.12230.08120.05780.10600.04740.09550.10330.1335
weight after (g)0.05410.05300.06500.13500.08090.05470.09270.04420.09480.09700.1233
difference %−7.04%−29.33%−8.84%10.38%−0.37%−5.36%−12.55%−6.75%−0.73%−6.10%−7.64%

Share and Cite

MDPI and ACS Style

Konieczny, J.; Loos, K. Green Polyurethanes from Renewable Isocyanates and Biobased White Dextrins. Polymers 2019, 11, 256. https://doi.org/10.3390/polym11020256

AMA Style

Konieczny J, Loos K. Green Polyurethanes from Renewable Isocyanates and Biobased White Dextrins. Polymers. 2019; 11(2):256. https://doi.org/10.3390/polym11020256

Chicago/Turabian Style

Konieczny, Jakob, and Katja Loos. 2019. "Green Polyurethanes from Renewable Isocyanates and Biobased White Dextrins" Polymers 11, no. 2: 256. https://doi.org/10.3390/polym11020256

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop