[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (72)

Search Parameters:
Keywords = stereocomplex

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 3498 KiB  
Article
The Effect of Polyethylene Glycol on the Non-Isothermal Crystallization of Poly(L-lactide) and Poly(D-lactide) Blends
by Panthima Phuangthong, Wenwei Li, Jun Shen, Mohammadreza Nofar, Patnarin Worajittiphon and Yottha Srithep
Polymers 2024, 16(15), 2129; https://doi.org/10.3390/polym16152129 - 26 Jul 2024
Viewed by 535
Abstract
The formation of polylactide stereocomplex (sc-PLA), involving the blending of poly(L-lactide) (PLLA) and poly(D-lactide) (PDLA), enhances PLA materials by making them stronger and more heat-resistant. This study investigated the competitive crystallization behavior of homocrystals (HCs) and stereocomplex crystals (SCs) in a 50/50 PLLA/PDLA [...] Read more.
The formation of polylactide stereocomplex (sc-PLA), involving the blending of poly(L-lactide) (PLLA) and poly(D-lactide) (PDLA), enhances PLA materials by making them stronger and more heat-resistant. This study investigated the competitive crystallization behavior of homocrystals (HCs) and stereocomplex crystals (SCs) in a 50/50 PLLA/PDLA blend with added polyethylene glycol (PEG). PEG, with molecular weights of 400 g/mol and 35,000 g/mol, was incorporated at concentrations ranging from 5% to 20% by weight. Differential scanning calorimetry (DSC) analysis revealed that PEG increased the crystallization temperature, promoted SC formation, and inhibited HC formation. PEG also acted as a plasticizer, lowering both melting and crystallization temperatures. The second heating DSC curve showed that the pure PLLA/PDLA blend had a 57.1% fraction of SC while adding 5% PEG with a molecular weight of 400 g/mol resulted in complete SC formation. In contrast, PEG with a molecular weight of 35,000 g/mol was less effective, allowing some HC formation. Additionally, PEG consistently promoted SC formation across various cooling rates (2, 5, 10, or 20 °C/min), demonstrating a robust influence under different conditions. Full article
(This article belongs to the Section Polymer Chemistry)
Show Figures

Figure 1

Figure 1
<p>Enantiomeric PLA homopolymers blend to form polylactide stereocomplex (sc-PLA).</p>
Full article ">Figure 2
<p>Schematic of the preparation process for fabricating PLLA/PDLA films with different PEG concentrations for non-isothermal crystallization kinetics analysis.</p>
Full article ">Figure 3
<p>DSC cooling curves of (<b>a</b>) PLLA/PDLA-G4 and (<b>b</b>) PLLA/PDLA-G350 at a cooling rate of 10 °C/min.</p>
Full article ">Figure 4
<p>DSC-derived parameters during cooling: (<b>a</b>) crystallization temperature (T<sub>c</sub>), (<b>b</b>) the difference between onset crystallization temperature and T<sub>c</sub> (T<sub>onset</sub> − T<sub>c</sub>).</p>
Full article ">Figure 5
<p>DSC second heating curves of (<b>a</b>) PLLA/PDLA-G4 and (<b>b</b>) PLLA/PDLA-G350 at a rate of 10 °C/min.</p>
Full article ">Figure 6
<p>The molecular structures of PLLA, PDLA, and PEG.</p>
Full article ">Figure 7
<p>DSC derived parameter during the second heating (<b>a</b>) stereocomplex melting temperature (T<sub>m,sc</sub>), (<b>b</b>) the crystallinity of HC (X<sub>c,hc)</sub>, and (<b>c</b>) the relative fraction of SC crystallites of PLA-G4 and PLA-G350 evaluated from DSC measurement (f<sub>sc,DSC</sub>).</p>
Full article ">Figure 8
<p>DSC cooling curves of PLLA/PDLA, PLLA/PDLA-G4-10, and PLLA/PDLA-G350-10 at various cooling rates (indicated on the curves).</p>
Full article ">Figure 9
<p>The effect of cooling rate on (<b>a</b>) T<sub>c</sub> and (<b>b</b>) T<sub>c</sub> − T<sub>onset</sub> of PLLA/PDLA, PLLA/PDLA-G4-10, and PLLA/PDLA-G350.</p>
Full article ">Figure 10
<p>DSC second heating curves at a rate of 10 °C/min for PLLA/PDLA, PLLA/PDLA-G4-10, and PLLA/PDLA-G350-10 at various cooling rates (indicated on the curves).</p>
Full article ">Figure 11
<p>The relative crystallinity (X<sub>T</sub>) versus crystallization temperature (T) for PLLA/PDLA, PLLA/PDLA-G4-10, and PLLA/PDLA-G350-10 at various cooling rates.</p>
Full article ">Figure 12
<p>The relative crystallinity (X<sub>t</sub>) versus crystallization time (t) for PLLA/PDLA, PLLA/PDLA-G4-10, and PLLA/PDLA-G350-10 at various cooling rates.</p>
Full article ">Figure 13
<p>Fitting curves of lg[−ln(1 − X<sub>t</sub>)] with respect to lgt for PLLA/PDLA, PLLA/PDLA-G4-10, and PLLA/PDLA-G350-10 at various cooling rates.</p>
Full article ">Figure 14
<p>Plots of ln D versus 1/T<sub>c</sub> using the Takhor equation.</p>
Full article ">
23 pages, 9760 KiB  
Article
Dual-Temperature/pH-Sensitive Hydrogels with Excellent Strength and Toughness Crosslinked Using Three Crosslinking Methods
by Jiaqi Wang, Wanying Yang, Yutong Li, Xuerong Ma, Yuxin Xie, Guangyan Zhou and Shouxin Liu
Gels 2024, 10(7), 480; https://doi.org/10.3390/gels10070480 - 19 Jul 2024
Viewed by 620
Abstract
Hydrogels are widely used as excellent drug carriers in the field of biomedicine. However, their application in medicine is limited by their poor mechanical properties and softness. To improve the mechanical properties of hydrogels, a novel triple-network amphiphilic hydrogel with three overlapping crosslinking [...] Read more.
Hydrogels are widely used as excellent drug carriers in the field of biomedicine. However, their application in medicine is limited by their poor mechanical properties and softness. To improve the mechanical properties of hydrogels, a novel triple-network amphiphilic hydrogel with three overlapping crosslinking methods using a one-pot free-radical polymerization was synthesized in this study. Temperature-sensitive and pH-sensitive monomers were incorporated into the hydrogel to confer stimulus responsiveness, making the hydrogel stimuli-responsive. The successful synthesis of the hydrogel was confirmed using techniques, such as proton nuclear magnetic resonance spectroscopy (1H NMR), Fourier-transform infrared spectroscopy (FT-IR), and X-ray diffraction (XRD). In order to compare and analyze the properties of physically crosslinked hydrogels, physically–chemically double-crosslinked hydrogels, and physically–chemically clicked triple-crosslinked hydrogels, various tests were conducted on the gels’ morphology, swelling behavior, thermal stability, mechanical properties, and drug loading capacity. The results indicate that the triple-crosslinked hydrogel maintains low swelling, high mechanical strength, and good thermal stability while not significantly compromising its drug delivery capability. Full article
(This article belongs to the Section Gel Processing and Engineering)
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H NMR diagram of HEMA-PLLA<sub>30</sub>.</p>
Full article ">Figure 2
<p>FT−IR spectra of the HEMA-PLLA<sub>30</sub> (<b>a</b>), HEMA-(PLLA-PDLA) (<b>b</b>), HEMA-(PLLA-PDLA)-N<sub>3</sub> (<b>c</b>), HEMA-(PLLA-PDLA)-alkyne (<b>d</b>), and gel3 (<b>e</b>).</p>
Full article ">Figure 3
<p>(<b>a</b>) XRD pattern of the HEMA-PLLA; (<b>b</b>) XRD pattern of the HEMA-(PLLA-PDLA).</p>
Full article ">Figure 4
<p>FT−IR spectra of gel1 (<b>a</b>); gel2 (<b>b</b>); gel3 (<b>c</b>).</p>
Full article ">Figure 5
<p>(<b>a</b>) XRD pattern of gel1; (<b>b</b>) XRD pattern of gel2; (<b>c</b>) XRD pattern of gel3.</p>
Full article ">Figure 6
<p>SEM of gel1 (<b>a</b>); gel2 (<b>b</b>); gel3 (<b>c</b>).</p>
Full article ">Figure 7
<p>(<b>a</b>) Swelling rate of gels at 25 °C, where n is the number of samples; (<b>b</b>) deswelling rate of gels at 37 °C, where n is the number of samples.</p>
Full article ">Figure 8
<p>(<b>a</b>) Swelling rate of gels in solution at 37 °C and pH = 5, where n is the number of samples; (<b>b</b>) deswelling rate of gels in solution at 37 °C and pH = 9, where n is the number of samples.</p>
Full article ">Figure 9
<p>Reversible swelling of gel1, gel2 and gel3, where n is the number of samples.</p>
Full article ">Figure 10
<p>Swelling rate of gels at THF, where n is the number of samples.</p>
Full article ">Figure 11
<p>Digital photographs of gel1 (<b>a</b>), gel2 (<b>b</b>), and gel3 (<b>c</b>) in dry state, fully swollen in secondary deionized water, and fully swollen in THF.</p>
Full article ">Figure 12
<p>(<b>a</b>) TG diagram of gels; (<b>b</b>) DTG diagram of gels.</p>
Full article ">Figure 13
<p>Energy storage modulus curve of gels.</p>
Full article ">Figure 14
<p>(<b>a</b>) Drug release rates of the gels at pH = 5.3, 37 °C; (<b>b</b>) drug release rates of the gels at pH = 7.4, 37 °C; (<b>c</b>) drug release rates of the gels at pH = 7.4, 40 °C.</p>
Full article ">Scheme 1
<p>Synthesis of stereocomplexes HEMA-(PLLA-PDLA)<sub>n</sub>.</p>
Full article ">Scheme 2
<p>Synthesis of HEMA-(PLLA-PDLA)-N<sub>3</sub>, HEMA-(PLLA-PDLA)-alkyne and triazole polymers.</p>
Full article ">Scheme 3
<p>Synthesis of hydrogels.</p>
Full article ">
20 pages, 9641 KiB  
Article
The Effect of Stereocomplexation and Crystallinity on the Degradation of Polylactide Nanoparticles
by Chuan Yin, Jenny Hemstedt, Karl Scheuer, Maja Struczyńska, Christine Weber, Ulrich S. Schubert, Jörg Bossert and Klaus D. Jandt
Nanomaterials 2024, 14(5), 440; https://doi.org/10.3390/nano14050440 - 28 Feb 2024
Viewed by 1106
Abstract
Polymeric nanoparticles (PNPs) are frequently researched and used in drug delivery. The degradation of PNPs is highly dependent on various properties, such as polymer chemical structure, size, crystallinity, and melting temperature. Hence, a precise understanding of PNP degradation behavior is essential for optimizing [...] Read more.
Polymeric nanoparticles (PNPs) are frequently researched and used in drug delivery. The degradation of PNPs is highly dependent on various properties, such as polymer chemical structure, size, crystallinity, and melting temperature. Hence, a precise understanding of PNP degradation behavior is essential for optimizing the system. This study focused on enzymatic hydrolysis as a degradation mechanism by investigation of the degradation of PNP with various crystallinities. The aliphatic polyester polylactide ([C3H4O2]n, PLA) was used as two chiral forms, poly l-lactide (PlLA) and poly d-lactide (PdLA), and formed a unique crystalline stereocomplex (SC). PNPs were prepared via a nanoprecipitation method. In order to further control the crystallinity and melting temperatures of the SC, the polymer poly(3-ethylglycolide) [C6H8O4]n (PEtGly) was synthesized. Our investigation shows that the PNP degradation can be controlled by various chemical structures, crystallinity and stereocomplexation. The influence of proteinase K on PNP degradation was also discussed in this research. AFM did not reveal any changes within the first 24 h but indicated accelerated degradation after 7 days when higher EtGly content was present, implying that lower crystallinity renders the particles more susceptible to hydrolysis. QCM-D exhibited reduced enzyme adsorption and a slower degradation rate in SC-PNPs with lower EtGly contents and higher crystallinities. A more in-depth analysis of the degradation process unveiled that QCM-D detected rapid degradation from the outset, whereas AFM exhibited delayed changes of degradation. The knowledge gained in this work is useful for the design and creation of advanced PNPs with enhanced structures and properties. Full article
(This article belongs to the Special Issue Functional Nanomaterials for Theranostic Applications)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of P<span class="html-small-caps">l</span>LA, P<span class="html-small-caps">d</span>LA, P(<span class="html-small-caps">l</span>LA-<span class="html-italic">stat</span>-EtGly), and P(<span class="html-small-caps">d</span>LA-<span class="html-italic">stat</span>-EtGly).</p>
Full article ">Figure 2
<p>Schematic illustration of the nanoprecipitation process (good solvent here refers to a solvent with a comparably higher evaporation rate than the poor solvent).</p>
Full article ">Figure 3
<p>Hydrodynamic radius R<sub>h</sub> (bars) and polydispersity index PDI (red dots) of the prepared PNPs with different EtGly amounts (<b>a</b>) and DLS curves (<b>b</b>) for all PNPs in suspension. The error bars indicate the standard deviation (SD) from the mean.</p>
Full article ">Figure 4
<p>Particle heights of the prepared PNPs with different EtGly amounts determined by AFM (<b>a</b>) and representative AFM images of the PNPs on functionalized Si substrates (<b>b</b>). The height scale on the right side and the lateral scale in the bottom right image (400 nm) apply to all images.</p>
Full article ">Figure 4 Cont.
<p>Particle heights of the prepared PNPs with different EtGly amounts determined by AFM (<b>a</b>) and representative AFM images of the PNPs on functionalized Si substrates (<b>b</b>). The height scale on the right side and the lateral scale in the bottom right image (400 nm) apply to all images.</p>
Full article ">Figure 5
<p>Particle heights of SC-PNP samples before degradation determined using AFM. Lines below bars indicate statistical differences in particle height in the group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Particle heights of the SC-PNPs determined using AFM after degradation compared with the measured values before degradation. Lines below bars indicate statistical differences in particle height before and after degradation (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Area density of SC-PNPs before and after degradation on all samples determined using AFM (<b>a</b>) and percentage change in area density of SC-PNPs during degradation on all samples (<b>b</b>).</p>
Full article ">Figure 7 Cont.
<p>Area density of SC-PNPs before and after degradation on all samples determined using AFM (<b>a</b>) and percentage change in area density of SC-PNPs during degradation on all samples (<b>b</b>).</p>
Full article ">Figure 8
<p>Particle heights (<b>left</b>) and area densities (<b>middle</b>) of the SC-PNPs determined via AFM before and after degradation for 7 days and the percentage change in area density (<b>right</b>) during degradation. Lines below bars indicate statistical differences in particle height in the group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>QCM-D measurement curves of SC-PNP degradation.</p>
Full article ">Figure 10
<p>PNP particle heights without stereocomplexation determined via AFM before and after the degradation. (<b>a</b>–<b>d</b>) P<span class="html-small-caps">d</span>LA-PNPs, (<b>e</b>–<b>h</b>) P<span class="html-small-caps">l</span>LA-PNPs. Lines below bars indicate statistical differences in particle height in the group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 11
<p>Area density of the PNPs without stereocomplexation before and after degradation. (<b>a</b>–<b>d</b>) P<span class="html-small-caps">d</span>LA-PNPs, (<b>e</b>–<b>h</b>) P<span class="html-small-caps">l</span>LA-PNPs.</p>
Full article ">Figure 12
<p>QCM-D measurement curves of the degradation of P<span class="html-small-caps">l</span>LA-PNPs.</p>
Full article ">Figure 13
<p>Elastic moduli of all SC-PNPs (<b>a</b>) and all PNPs without stereocomplexation (<b>b</b>) determined using FDCs. Lines below bars indicate statistical differences in elastic moduli in the group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
13 pages, 3729 KiB  
Article
Gelation upon the Mixing of Amphiphilic Graft and Triblock Copolymers Containing Enantiomeric Polylactide Segments through Stereocomplex Formation
by Yuichi Ohya, Yasuyuki Yoshida, Taiki Kumagae and Akinori Kuzuya
Gels 2024, 10(2), 139; https://doi.org/10.3390/gels10020139 - 9 Feb 2024
Viewed by 1623
Abstract
Biodegradable injectable polymer (IP) systems that form hydrogels in situ when injected into the body have considerable potential as medical materials. In this paper, we report a new two-solution mixed biodegradable IP system that utilizes the stereocomplex (SC) formation of poly(l-lactide) [...] Read more.
Biodegradable injectable polymer (IP) systems that form hydrogels in situ when injected into the body have considerable potential as medical materials. In this paper, we report a new two-solution mixed biodegradable IP system that utilizes the stereocomplex (SC) formation of poly(l-lactide) (PLLA) and poly(d-lactide) (PDLA). We synthesized triblock copolymers of PLLA and poly(ethylene glycol), PLLA-b-PEG-b-PLLA (tri-L), and a graft copolymer of dextran (Dex) attached to a PDLA-b-PEG diblock copolymer, Dex-g-(PDLA-b-PEG) (gb-D). We found that a hydrogel can be obtained by mixing gb-D solution and tri-L solution via SC formation. Although it is already known that graft copolymers attached to enantiomeric PLLA and PDLA chains can form an SC hydrogel upon mixing, we revealed that hydrogels can also be formed by a combination of graft and triblock copolymers. In this system (graft vs. triblock), the gelation time was shorter, within 1 min, and the physical strength of the resulting hydrogel (G′ > 100 Pa) was higher than when graft copolymers were mixed. Triblock copolymers form micelles (16 nm in diameter) in aqueous solutions and hydrophobic drugs can be easily encapsulated in micelles. In contrast, graft copolymers have the advantage that their molecular weight can be set high, contributing to improved mechanical strength of the obtained hydrogel. Various biologically active polymers can be used as the main chains of graft copolymers, and chemical modification using the remaining functional side chain groups is also easy. Therefore, the developed mixing system with a graft vs. triblock combination can be applied to medical materials as a highly convenient, physically cross-linked IP system. Full article
(This article belongs to the Special Issue Recent Advances in Thermoreversible Gelation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustration of the polymers, Dex-<span class="html-italic">g</span>-(PLLA-<span class="html-italic">b</span>-PEG) (<span class="html-italic">gb-</span>L), Dex-<span class="html-italic">g</span>-(PDLA-<span class="html-italic">b</span>-PEG) (<span class="html-italic">gb-</span>D), and PLLA-<span class="html-italic">b</span>-PEG-<span class="html-italic">b</span>-PLLA (<span class="html-italic">tri</span>-L) used in this study and hydrogel formation by mixing of <span class="html-italic">gb-</span>D and <span class="html-italic">tri-</span>L.</p>
Full article ">Figure 2
<p><sup>1</sup>H-NMR spectra of (<b>A</b>) PLLA-<span class="html-italic">b</span>-PEG-<span class="html-italic">b</span>-PLLA triblock copolymer (<span class="html-italic">tri</span>-L) (solvent: CDCl<sub>3</sub>), (<b>B</b>) MeO-PEG-<span class="html-italic">b</span>-PDLA diblock copolymer (<span class="html-italic">b</span>-D) (solvent: CDCl<sub>3</sub>), and (<b>C</b>) Dex-<span class="html-italic">g</span>-(PDLA-<span class="html-italic">b</span>-PEG) (<span class="html-italic">gb</span>-D) (solvent: D<sub>2</sub>O/NaOD). The lowercase letters a–d for (<b>A</b>), a–f for (<b>B</b>) and a–h for (<b>C</b>) indicate the assignments of each peak.</p>
Full article ">Figure 3
<p>Photographs of the typical examples of polymer solutions before mixing and after gelation. The final polymer concentrations are shown in <a href="#gels-10-00139-t003" class="html-table">Table 3</a>.</p>
Full article ">Figure 4
<p>Effects of L-lactide unit/D-lactide unit ratio (L/D) on (<b>A</b>) gelation time and (<b>B</b>) storage modulus (G′) 24 h after mixing at 37 °C for the combination of PLLA-<span class="html-italic">b</span>-PEG-<span class="html-italic">b</span>-PLLA (<span class="html-italic">tri</span>-L) and Dex-<span class="html-italic">g</span>-(PDLA-<span class="html-italic">b</span>-PEG) (<span class="html-italic">gb-</span>D).</p>
Full article ">Figure 5
<p>Time course of storage modulus (G′, closed symbols) and loss modulus (G″, open symbols) for the mixture of PLLA-<span class="html-italic">b</span>-PEG-<span class="html-italic">b</span>-PLLA (<span class="html-italic">tri</span>-L) and Dex-<span class="html-italic">g</span>-(PDLA-<span class="html-italic">b</span>-PEG) (<span class="html-italic">gb-D</span>) (L/D = 2.0) at 37 °C (red circles) and 25 °C (blue triangles).</p>
Full article ">Figure 6
<p>Wide-angle X-ray diffraction (WAXD) spectra for the mixture of <span class="html-italic">tri-</span>L and <span class="html-italic">gb-</span>D (L/D = 1.0 and 5.0) after freeze-drying and intact dextran.</p>
Full article ">Figure 7
<p>Results of dynamic light scattering (DLS) measurement for <span class="html-italic">tri</span>-L solution (<b>a</b>) and <span class="html-italic">gb</span>-D so lution (<b>b</b>) in water at 25 °C expressed as number distribution. Polymer concentration = 0.2 wt%.</p>
Full article ">Scheme 1
<p>Synthesis of PLLA-<span class="html-italic">b</span>-PEG-<span class="html-italic">b</span>-PLLA (<span class="html-italic">tri</span>-L).</p>
Full article ">Scheme 2
<p>Synthesis of Dex-<span class="html-italic">g</span>-PLLA-<span class="html-italic">b</span>-PEG and Dex-<span class="html-italic">g</span>-PDLA-<span class="html-italic">b</span>-PEG.</p>
Full article ">
14 pages, 4188 KiB  
Article
Properties of Stereocomplex PLA for Melt Spinning
by Boris Marx, Lars Bostan, Axel S. Herrmann, Laura Boskamp and Katharina Koschek
Polymers 2023, 15(23), 4510; https://doi.org/10.3390/polym15234510 - 23 Nov 2023
Viewed by 1231
Abstract
Fibers made from biopolymers are one solution for conserving both resources and the environment. However, these fibers currently have limited strengths, which limit their use for textile applications. In this paper, a biopolymer stereocomplex poly(-lactide) (scPLA) formation on a technical scale of high-molecular-weight [...] Read more.
Fibers made from biopolymers are one solution for conserving both resources and the environment. However, these fibers currently have limited strengths, which limit their use for textile applications. In this paper, a biopolymer stereocomplex poly(-lactide) (scPLA) formation on a technical scale of high-molecular-weight poly(D-lactide) (PDLA) and poly(L-lactide) (PLLA) is presented. This scPLA material is the basis for further research to develop scPLA yarns in melt spinning with technical strengths for technical application. scPLA is compared with standard and commercially available semi-crystalline PLA for the production of fibers in melt spinning (msPLA) with textile strengths. Differential scanning calorimetry (DSC) gives a degree of crystallization of 59.7% for scPLA and 47.0% for msPLA. X-ray diffraction (XRD) confirms the pure stereocomplex crystal structure for scPLA and semi-crystallinity for msPLA. scPLA and msPLA are also compared regarding their processing properties (rheology) in melt spinning. While complex viscosity of scPLA is much lower compared to msPLA, both materials show similar viscoelastic behavior. Thermal gravimetric analysis (TGA) shows the influence of the molecular weight on the thermal stability, whereas essentially the crystallinity influences the biodegradability of the PLA materials. Full article
(This article belongs to the Special Issue Preparation and Applications of Biodegradable Polymer Materials)
Show Figures

Figure 1

Figure 1
<p>scPLA precipitation (<b>a</b>) and SEM image of scPLA powder (<b>b</b>).</p>
Full article ">Figure 2
<p>DSC thermograms (<b>a</b>) and XRD patterns (<b>b</b>) of msPLA and scPLA.</p>
Full article ">Figure 3
<p>Complex viscosity (<b>a</b>) and loss factor (<b>b</b>) of msPLA and scPLA.</p>
Full article ">Figure 4
<p>TGA results of PLLA, PDLA, msPLA and scPLA in granular/powder form.</p>
Full article ">Figure 5
<p>Photos of PLLA, PDLA, msPLA and scPLA samples stored in the soil after 0, 63 and 94 days. All samples were rinsed with distilled water.</p>
Full article ">Figure 6
<p>Weight loss after 63 days and 94 days for PLLA, PDLA, msPLA and scPLA.</p>
Full article ">Figure 7
<p>Deriv. weight curves of PLLA, PDLA, msPLA and scPLA after different decomposition times.</p>
Full article ">
19 pages, 6754 KiB  
Article
Stereo-Complex and Click-Chemical Bicrosslinked Amphiphilic Network Gels with Temperature/pH Response
by Wanying Yang, Jiaqi Wang, Lingjiang Jia, Jingyi Li and Shouxin Liu
Gels 2023, 9(8), 647; https://doi.org/10.3390/gels9080647 - 11 Aug 2023
Cited by 3 | Viewed by 1163
Abstract
Stimulus-responsive hydrogels have been widely used in the field of drug delivery because of their three-dimensional pore size and the ability to change the drug release rate with the change in external environment. In this paper, the temperature-sensitive monomer 2-methyl-2-acrylate-2-(2-methoxyethoxy-ethyl) ethyl ester (MEO [...] Read more.
Stimulus-responsive hydrogels have been widely used in the field of drug delivery because of their three-dimensional pore size and the ability to change the drug release rate with the change in external environment. In this paper, the temperature-sensitive monomer 2-methyl-2-acrylate-2-(2-methoxyethoxy-ethyl) ethyl ester (MEO2MA) and oligoethylene glycol methyl ether methacrylate (OEGMA) as well as the pH-sensitive monomer N,N-Diethylaminoethyl methacrylate (DEAEMA) were used to make the gel with temperature and pH response. Four kinds of physicochemical double-crosslinked amphiphilic co-network gels with different polymerization degrees were prepared by the one-pot method using the stereocomplex between polylactic acid as physical crosslinking and click chemistry as chemical crosslinking. By testing morphology, swelling, thermal stability and mechanical properties, the properties of the four hydrogels were compared. Finally, the drug release rate of the four gels was tested by UV–Vis spectrophotometer. It was found that the synthetic hydrogels had a good drug release rate and targeting, and had great application prospect in drug delivery. Full article
(This article belongs to the Special Issue Hydrogels in Action: Self-Assembly, Responsivity and Sensing)
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H NMR diagram of HEMA−PLLA<sub>10</sub>.</p>
Full article ">Figure 2
<p>(<b>a</b>) <sup>1</sup>H NMR diagram of HEMA−PLLA<sub>10</sub>; (<b>b</b>) <sup>1</sup>H NMR diagram of HEMA−PLLA<sub>20</sub>; (<b>c</b>) <sup>1</sup>H NMR diagram of HEMA−PLLA<sub>30</sub>; (<b>d</b>) <sup>1</sup>H NMR diagram of HEMA−PLLA<sub>40</sub>.</p>
Full article ">Figure 3
<p>FT−IR spectrum of the macromolecular monomer.</p>
Full article ">Figure 4
<p>(a) FT−IR spectrum of the macromolecular monomer; (b) FT−IR spectrum of HEMA−PDLA−alkyne; (c) FT−IR spectrum ofHEMA−PLLA−N<sub>3</sub>.</p>
Full article ">Figure 5
<p>(a) FT−IR spectrum of HEMA−PDLA−alkyne; (b) FT−IR spectrum of HEMA−PLLA−N<sub>3</sub>; (c) FT−IR spectrum of gel.</p>
Full article ">Figure 6
<p>(a) XRD pattern of the gel; (b) XRD pattern of the macromolecular monomer.</p>
Full article ">Figure 7
<p>SEM of gel 1 (<b>a</b>); gel 2 (<b>b</b>); gel 3 (<b>c</b>); gel 4 (<b>d</b>).</p>
Full article ">Figure 8
<p>(<b>a</b>) Swelling rate of gels in distilled water at 25 °C; (<b>b</b>) swelling rate of gels at THF.</p>
Full article ">Figure 9
<p>(<b>a</b>) Digital photo of the gel 1 completely swelling in distilled water; (<b>b</b>) digital photo of the gel 1 completely swelling in THF; (<b>c</b>) digital photo of dry gel 1.</p>
Full article ">Figure 10
<p>Deswelling rate of gels in distilled water at 37 °C.</p>
Full article ">Figure 11
<p>Reversible swelling of gel 1 and gel 4.</p>
Full article ">Figure 12
<p>(<b>a</b>) Swelling rate of gels in solution at 37 °C and pH = 5; (<b>b</b>) deswelling rate of gels in solution at 37 °C and pH = 9.</p>
Full article ">Figure 13
<p>Molecular structural formulae of DEAEMA at different pH values.</p>
Full article ">Figure 14
<p>Thermal analysis diagram of gels.</p>
Full article ">Figure 15
<p>DTG diagram of gels.</p>
Full article ">Figure 16
<p>Energy storage modulus curve of gels.</p>
Full article ">Figure 17
<p>(<b>a</b>) Drug release rate of the gel at pH = 5.2, 37 °C; (<b>b</b>) drug release rates of gel 1 and gel 4 at pH = 7.4, 37 °C and pH = 5.2, 37 °C.</p>
Full article ">Scheme 1
<p>Synthesis of macromolecular monomer HEMA-PLLA<sub>n</sub> and HEMA-PDLA<sub>n</sub>.</p>
Full article ">Scheme 2
<p>Synthesis of HEMA-PLLA-N<sub>3</sub> and HEMA-PDLA-alkyne.</p>
Full article ">Scheme 3
<p>Synthesis of stereocomplex and click-chemical double-crosslinked gels.</p>
Full article ">
13 pages, 3306 KiB  
Article
Influence of Molecular Weight on the Enzymatic Degradation of PLA Isomer Blends by a Langmuir System
by Donghyeok Im, Vishal Gavande, Hak Yong Lee and Won-Ki Lee
Materials 2023, 16(14), 5087; https://doi.org/10.3390/ma16145087 - 19 Jul 2023
Cited by 2 | Viewed by 1001
Abstract
Polylactides (PLAs) and lactide copolymers are biodegradable, compostable, and derived from renewable resources, offering a sustainable alternative to petroleum-based synthetic polymers owing to their advantages of comparable mechanical properties with commodity plastics and biodegradability. Their hydrolytic stability and thermal properties can affect their [...] Read more.
Polylactides (PLAs) and lactide copolymers are biodegradable, compostable, and derived from renewable resources, offering a sustainable alternative to petroleum-based synthetic polymers owing to their advantages of comparable mechanical properties with commodity plastics and biodegradability. Their hydrolytic stability and thermal properties can affect their potential for long-lasting applications. However, stereocomplex crystallization is a robust method between isomer PLAs that allows significant amelioration in copolymer properties, such as thermal stability, mechanical properties, and biocompatibility, through substantial intermolecular interactions amid l-lactyl and d-lactyl sequences, which have been the key approach to initial degradation rate and further PLA applications. It was demonstrated that the essential parameters affecting stereocomplexation are the mixing ratio and the chain length of each unit sequence. This study deals with the molecular weight, one of the specific interactions between isomers of PLAs. A solution polymerization method was applied to control molecular weight and chain architecture. The stereocomplexation was monitored with DSC. It was confirmed that the lower molecular weight polymer showed a higher degradation rate, as a hydrolyzed fragment having a molecular weight below a certain length dissolves into the water. To systematically explore the critical contribution of molecular weights, the Langmuir system was used to observe the stereocomplexation effect and the overall degradation rate. Full article
(This article belongs to the Special Issue Bio-Based Materials and Their Environmental Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the degradation kinetics of the Langmuir monolayer.</p>
Full article ">Figure 2
<p>Schematic diagram of stereocomplexation according to molecular weight.</p>
Full article ">Figure 3
<p>DSC curves of isomer PLA (<b>A</b>) and PLLA/PDLA blend films (<b>B</b>).</p>
Full article ">Figure 4
<p>X<sub>sc</sub> and T<sub>sc</sub> of PLA stereocomplexes with different molecular weights, where M<sub>n</sub> of PDLA was fixed to 90 k (X<sub>sc</sub> and T<sub>sc</sub> of LLStereo indicated by <b>★</b>).</p>
Full article ">Figure 5
<p>π-A isotherms of PLLA and stereocomplexed monolayers at pH 8.6.</p>
Full article ">Figure 6
<p>π-A isotherms of (<b>A</b>) LLStereo and (<b>B</b>) HHStereo at pH 8.6: a: before degradation; b: after enzymatic degradation.</p>
Full article ">Figure 7
<p>Area ratio vs. time of PLA homopolymer (<b>A</b>) and stereocomplex (<b>B</b>) at 4 mN/m on the subphase of pH 8.6 with proteinase K.</p>
Full article ">Figure 8
<p>SEM images of PLLA/PDLA blends before (left) and after (right) enzymatic degradation (6 h); (<b>A</b>): LLStereo, (<b>B</b>): LHStereo, and (<b>C</b>): HHStereo.</p>
Full article ">
12 pages, 4077 KiB  
Article
Preparation and Properties of Physical Gel on Medical Titanium Alloy Surface
by Yu Fu, Qingrong Wu, Wanying Yang, Jiaqi Wang, Zechen Liu, Hao Shi and Shouxin Liu
Gels 2023, 9(7), 558; https://doi.org/10.3390/gels9070558 - 8 Jul 2023
Viewed by 1232
Abstract
Medical titanium alloy Ti-6Al-4V (TC4) has been widely used in the medical field, especially in human tissue repair. However, TC4 has some shortcomings, which may cause problems with biocompatibility and mechanical compatibility in direct contact with the human body. To solve this problem, [...] Read more.
Medical titanium alloy Ti-6Al-4V (TC4) has been widely used in the medical field, especially in human tissue repair. However, TC4 has some shortcomings, which may cause problems with biocompatibility and mechanical compatibility in direct contact with the human body. To solve this problem, physical gels are formed on the surface of TC4, and the storage modulus of the formed physical gel matches that of the human soft tissue. 2-bromoisobutyryl bromide (BIBB) and dopamine (DA) were used to form initiators on the surface of hydroxylated medical titanium alloy. Different initiators were formed by changing the ratio of BIBB and DA, and the optimal one was selected for subsequent reactions. Under the action of the catalyst, L-lactide and D-lactide were ring-opened polymerized with hydroxyethyl methacrylate (HEMA), respectively, to form macromolecular monomers HEMA-PLLA29 and HEMA-PDLA29 with a polymerization degree of 29. The two macromolecular monomers were stereo-complexed by ultrasound to form HEMA-stereocomplex polylactic acid (HEMA-scPLA29). Based on two monomers, 2-(2-methoxyethoxy) ethyl methacrylate (MEO2MA) and oligo (ethylene oxide) methacrylate (OEGMA), and the physical crosslinking agent HEMA-scPLA29, physical gels are formed on the surface of TC4 attached to the initiator via Atom Transfer Radical Addition Reaction (ATRP) technology. The hydrogels on the surface of titanium alloy were characterized and analyzed by a series of instruments. The results showed that the storage modulus of physical glue was within the range of the energy storage modulus of human soft tissue, which was conducive to improving the mechanical compatibility of titanium alloy and human soft tissue. Full article
(This article belongs to the Special Issue Advance in Composite Gels (2nd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The <sup>1</sup>H NMR spectrum of HEMA-PLLA<sub>29</sub>.</p>
Full article ">Figure 2
<p>Infrared spectra of initiator TC4−OH@DA/BIBB<sub>n</sub> (the molar ratio of DA/BIBB was 1:0.5, 1:1, and 1:1.5, respectively).</p>
Full article ">Figure 3
<p>Microstructure of initiators TC4-OH@DA/BIBB<sub>0.5</sub> (<b>a</b>), TC4-OH@DA/BIBB<sub>1</sub> (<b>b</b>), and TC4-OH@DA/BIBB<sub>1.5</sub> (<b>c</b>) at the same multiples. TC4-OH@DA/BIBB<sub>0.5</sub> (<b>d</b>), TC4-OH@DA/BIBB<sub>1</sub> (<b>e</b>), and TC4-OH@DA/BIBB<sub>1.5</sub> (<b>f</b>) were the microstructure of initiators at the same multiples [enlarge the local parts of (<b>a</b>–<b>c</b>)].</p>
Full article ">Figure 4
<p>XRD patterns of stereocomplex macromolecular monomers and physical hydrogels PGel.</p>
Full article ">Figure 5
<p>SEM images of physical hydrogels PGel1 (<b>a</b>), PGel2 (<b>b</b>), PGel3 (<b>c</b>).</p>
Full article ">Figure 6
<p>The variation curves of physical hydrogels PGel storage modulus with frequency.</p>
Full article ">Figure 7
<p>The loss tangents of physical hydrogels PGel at a frequency of 10 Hz.</p>
Full article ">Scheme 1
<p>Synthesis of macromolecular monomer HEMA-PDLA<sub>29</sub>(PLLA<sub>29</sub>).</p>
Full article ">Scheme 2
<p>Schematic diagram of the synthesis of the physical crosslinking agent HEMA-scPLA<sub>29</sub>.</p>
Full article ">Scheme 3
<p>Schematic diagram of the synthesis of physical hydrogels PGel on the surface of medical titanium alloys.</p>
Full article ">
22 pages, 7315 KiB  
Review
Development of Stereocomplex Polylactide Nanocomposites as an Advanced Class of Biomaterials—A Review
by Muhammad Samsuri and Purba Purnama
Polymers 2023, 15(12), 2730; https://doi.org/10.3390/polym15122730 - 19 Jun 2023
Cited by 3 | Viewed by 1520
Abstract
This review paper analyzes the development of advanced class polylactide (PLA) materials through a combination of stereocomplexation and nanocomposites approaches. The similarities in these approaches provide the opportunity to generate an advanced stereocomplex PLA nanocomposite (stereo-nano PLA) material with various beneficial properties. As [...] Read more.
This review paper analyzes the development of advanced class polylactide (PLA) materials through a combination of stereocomplexation and nanocomposites approaches. The similarities in these approaches provide the opportunity to generate an advanced stereocomplex PLA nanocomposite (stereo-nano PLA) material with various beneficial properties. As a potential “green” polymer with tunable characteristics (e.g., modifiable molecular structure and organic–inorganic miscibility), stereo-nano PLA could be used for various advanced applications. The molecular structure modification of PLA homopolymers and nanoparticles in stereo-nano PLA materials enables us to encounter stereocomplexation and nanocomposites constraints. The hydrogen bonding of D- and L-lactide fragments aids in the formation of stereococomplex crystallites, while the hetero-nucleation capabilities of nanofillers result in a synergism that improves the physical, thermal, and mechanical properties of materials, including stereocomplex memory (melt stability) and nanoparticle dispersion. The special properties of selected nanoparticles also allow the production of stereo-nano PLA materials with distinctive characteristics, such as electrical conductivity, anti-inflammatory, and anti-bacterial properties. The D- and L-lactide chains in PLA copolymers provide self-assembly capabilities to form stable nanocarrier micelles for encapsulating nanoparticles. This development of advanced stereo-nano PLA with biodegradability, biocompatibility, and tunability properties shows potential for use in wider and advanced applications as a high-performance material, in engineering field, electronic, medical device, biomedical, diagnosis, and therapeutic applications. Full article
(This article belongs to the Special Issue Biodegradable Polymers: Synthesis, Characterization and Applications)
Show Figures

Figure 1

Figure 1
<p>Crystal structure of s-PLA. (<b>A</b>) Structural model of s-PLA crystal from PLLA and PDLA. (<b>B</b>) Molecular arrangement projected on the plan normal to the chain axis. Reproduced from [<a href="#B5-polymers-15-02730" class="html-bibr">5</a>], with permission from Taylor &amp; Francis, 2022.</p>
Full article ">Figure 2
<p>Properties of PLLA, PDLA and s-PLA (PLLA/PDLA) materials; mechanical properties as function of molecular weight. (<b>a</b>) Tensile strength (<b>b</b>) Young’s modulus (<b>c</b>) Elongation at break and thermal properties [<a href="#B14-polymers-15-02730" class="html-bibr">14</a>] (copyright and permission, Elsevier 2023). (<b>d</b>) Melting temperature. (<b>e</b>) Thermal degradation temperature [<a href="#B22-polymers-15-02730" class="html-bibr">22</a>] (Copyright and permission, American Chemical Society 2023). Reproduced from [<a href="#B22-polymers-15-02730" class="html-bibr">22</a>], with permission from American Chemical Society, 2023.</p>
Full article ">Figure 3
<p>Preparation methods of PLA Nanocomposites.</p>
Full article ">Figure 4
<p>Material combination of stereo-nano PLA.</p>
Full article ">Figure 5
<p>Polarized optical microscope images of spherulites of neat PLLA and various PLLA blends after isothermal crystallization at 135 °C for 6 h: (<b>a</b>) neat PLLA, (<b>b</b>) 5% lignin (PLLA with 5% lignin content), (<b>c</b>) 5% LGDPD<sub>L</sub> (PLLA with 5% lignin-grafted PDLA; 1100 Da block length), (<b>d</b>) 5% LGDPD<sub>H</sub> (PLLA with 5% lignin-grafted PDLA; 2000 Da block length), (<b>e</b>) 5% 2a (PLLA with 5% 2-armed PDLA), (<b>f</b>) 5% 4a (PLLA with 5% 4-armed PDLA). Reproduced from [<a href="#B130-polymers-15-02730" class="html-bibr">130</a>], with permission from Elsevier, 2022.</p>
Full article ">Figure 6
<p>(<b>a</b>) Thermal degradation profile of lignin, PLA homoplymer, s-PLA and stereo-nano-PLA (3% s-PLA-Lig and 5% s-PLA-Lig) [<a href="#B98-polymers-15-02730" class="html-bibr">98</a>]. (<b>b</b>) Thermal degradation profile of PLA homopolymer, PLA nanocomposite with 3% of clays (3% nano PLA), s-PLA, and stereo-nano PLA with 3% of clays (3% s-PLA_Clay). Reproduced from [<a href="#B85-polymers-15-02730" class="html-bibr">85</a>], with permission from Springer Nature, 2022.</p>
Full article ">Figure 7
<p>The DSC thermograms of stereo-nano PLA from MWCNT-g-PLLA and MWCNT-g-PDLA. (<b>a</b>) Precipitated s-PLA-21 and (<b>b</b>) thin-film s-PLA-21. The sample was heated from 0 to 250 °C, cooled down to 0 °C, kept for 1 min, and heated to 250 °C with a heating and cooling rate of 10 °C; the solid line represents the first heating run, and the dashed line represents the second heating run. Reproduced from [<a href="#B93-polymers-15-02730" class="html-bibr">93</a>], with permission from American Chemical Society, 2023.</p>
Full article ">Scheme 1
<p>Schematic illustration of stereo-nano PLA micelles with encapsulating stability supported by stereocomplex formation. Reproduced from [<a href="#B120-polymers-15-02730" class="html-bibr">120</a>], with permission from American Chemical Society, 2023.</p>
Full article ">
16 pages, 7063 KiB  
Article
Super-Tough and Biodegradable Poly(lactide-co-glycolide) (PLGA) Transparent Thin Films Toughened by Star-Shaped PCL-b-PDLA Plasticizers
by Jieun Jeong, Sangsoo Yoon, Xin Yang and Young Jun Kim
Polymers 2023, 15(12), 2617; https://doi.org/10.3390/polym15122617 - 8 Jun 2023
Cited by 3 | Viewed by 1992
Abstract
To obtain fully degradable and super-tough poly(lactide-co-glycolide) (PLGA) blends, biodegradable star-shaped PCL-b-PDLA plasticizers were synthesized using natural originated xylitol as initiator. These plasticizers were blended with PLGA to prepare transparent thin films. Effects of added star-shaped PCL-b-PDLA plasticizers on [...] Read more.
To obtain fully degradable and super-tough poly(lactide-co-glycolide) (PLGA) blends, biodegradable star-shaped PCL-b-PDLA plasticizers were synthesized using natural originated xylitol as initiator. These plasticizers were blended with PLGA to prepare transparent thin films. Effects of added star-shaped PCL-b-PDLA plasticizers on mechanical, morphological, and thermodynamic properties of PLGA/star-shaped PCL-b-PDLA blends were investigated. The stereocomplexation strong cross-linked network between PLLA segment and PDLA segment effectively enhanced interfacial adhesion between star-shaped PCL-b-PDLA plasticizers and PLGA matrix. With only 0.5 wt% addition of star-shaped PCL-b-PDLA (Mn = 5000 g/mol), elongation at break of the PLGA blend reached approximately 248%, without any considerable sacrifice over excellent mechanical strength and modulus of PLGA. Full article
(This article belongs to the Special Issue Biobased and Biodegradable Polymer Blends and Composites)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular weight information of PLGA.</p>
Full article ">Figure 2
<p><sup>1</sup>H-NMR analysis of PLGA copolymer.</p>
Full article ">Figure 3
<p><sup>1</sup>H-NMR analysis of star-shaped PCL-<span class="html-italic">b</span>-PDLA.</p>
Full article ">Figure 4
<p>FT−IR spectra of (<b>a</b>) star-shaped PCL-<span class="html-italic">b</span>-PDLA, PCL, and PDLA; (<b>b</b>) expanded spectra of ester carbonyl (OC=O) stretch bands of star-shaped PCL-<span class="html-italic">b</span>-PDLA copolymer and homo PCL; (<b>c</b>) expanded spectra of C-OH vibrations of star-shaped PCL-<span class="html-italic">b</span>-PDLA and homo PCL.</p>
Full article ">Figure 5
<p>XRD pattern of PLGA, PLGA/star-shaped PCL-<span class="html-italic">b</span>-PLA and star-shaped PCL-<span class="html-italic">b</span>-PDLA plasticizer.</p>
Full article ">Figure 6
<p>Mechanical properites of PLGA blended with star-shaped PCL-<span class="html-italic">b</span>-PDLA: (<b>a</b>) neat PLGA, PLLA and PDLA; (<b>b</b>) PLGA blended with PCL-<span class="html-italic">b</span>-PDLA Mw 5000; (<b>c</b>) PLGA blended with PCL-<span class="html-italic">b</span>-PDLA Mw 10,000; (<b>d</b>) PLGA blended with PCL-<span class="html-italic">b</span>-PDLA Mw 15,000.</p>
Full article ">Figure 7
<p>Thermal properties of PLGA blended with star-shaped PCL-<span class="html-italic">b</span>-PDLA films: (<b>a</b>) Differential Scanning Calorimetry analysis; (<b>b</b>) Thermogravimetric analysis.</p>
Full article ">Figure 8
<p>Surface morphology properties of PLGA blended with star-shaped PCL-<span class="html-italic">b</span>-PDLA: (<b>a</b>) Surface morphology of neat PLGA; (<b>b</b>,<b>c</b>) Surface morphology of PLGA/star-shaped PCL-<span class="html-italic">b</span>-PDLA.</p>
Full article ">Figure 9
<p>Cross-sectional images of PLGA and PLGA blended with star-shaped PCL-<span class="html-italic">b</span>-PDLA: (<b>a</b>) Cross-sectional images of neat PLGA; (<b>b</b>) Cross sectional image of PLGA blended with PCL-<span class="html-italic">b</span>-PDLA Mw 5000; (<b>c</b>) Cross sectional image of PLGA blended with PCL-<span class="html-italic">b</span>-PDLA Mw 10,000; (<b>d</b>) Cross sectional image of PLGA blended with PCL-<span class="html-italic">b</span>-PDLA Mw 15,000.</p>
Full article ">Figure 10
<p>Cross-sectional images of tensile-fracture and Surface image of whitening area in tensile fracture: (<b>a</b>) Cross-sectional image of PLGA; (<b>b</b>) Cross-sectional image of PLGA/star-shaped PCL-<span class="html-italic">b</span>-PDLA 10K 2%; (<b>c</b>) Surface image of whitening area of PLGA/star-shaped PCL-<span class="html-italic">b</span>-PDLA 10K 2%; (<b>d</b>) Magnified picture of the area highlighted by the dotted line of (<b>c</b>).</p>
Full article ">Figure 11
<p>Transparency of PLGA and PLGA/Star-shaped PCL-<span class="html-italic">b</span>-PDLA.</p>
Full article ">Scheme 1
<p>Synthesis route of PLGA.</p>
Full article ">Scheme 2
<p>Synthesis route of star-shaped poly(ε-caprolactone-co-D-lactide) block copolymers.</p>
Full article ">
16 pages, 6291 KiB  
Article
Toughening Polylactide Stereocomplex by Injection Molding with Thermoplastic Starch and Chain Extender
by Yottha Srithep, Dutchanee Pholharn, Patnarin Worajittiphon, Keartisak Sriprateep, Onpreeya Veang-in and John Morris
Polymers 2023, 15(9), 2055; https://doi.org/10.3390/polym15092055 - 26 Apr 2023
Cited by 4 | Viewed by 1544
Abstract
The high cost, low heat resistance, and brittleness of poly(L-lactide) (PLLA) is a significant drawback that inhibits its diffusion into many industrial applications. These weaknesses were solved by forming a polylactide stereocomplex (ST) and blending it with thermoplastic starch (TPS). We blended poly [...] Read more.
The high cost, low heat resistance, and brittleness of poly(L-lactide) (PLLA) is a significant drawback that inhibits its diffusion into many industrial applications. These weaknesses were solved by forming a polylactide stereocomplex (ST) and blending it with thermoplastic starch (TPS). We blended poly (L-lactide)(PLLA), up to 30% thermoplastic starch, and a chain extender (2%) in an internal mixer, which was then hand-mixed with poly (D-lactide)(PDLA) and injection molded to form specimens, in order to study mechanical, thermal, and crystallization behavior. Differential scanning calorimetry (DSC) and wide-angle X-ray diffraction (XRD) demonstrated that the stereocomplex structures were still formed despite the added TPS and showed melting points ~55 °C higher than neat PLLA. Furthermore, stereocomplex crystallinity decreased with the increased TPS content. Dynamic mechanical analysis revealed that ST improved PLLA heat resistance, and tensile testing suggested that the TPS improved the elongation-at-break of ST. Moreover, the chain extender reduced the degradation of ST/TPS blends and generally improved ST/TPS composites’ mechanical properties. Full article
(This article belongs to the Special Issue Processing and Application of Bio-Based Polymeric Compounds)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Appearance of hand-mixed PLLA and PDLA, injection molded at 180 °C and (<b>b</b>) injection-molded samples formed at 225 °C with varying amounts of TPS and 2% chain extender.</p>
Full article ">Figure 2
<p>DSC melting curves for PLLA, ST, TPS, and ST/TPS/CE blends.</p>
Full article ">Figure 3
<p>XRD profiles of PLA and ST/starch blends without CE and with 2% CE content.</p>
Full article ">Figure 4
<p>Nonisothermal measurement of the percentage of remaining weight of PLLA, TPS, ST/TPS/CE at a constant 10 °C/min heating rate.</p>
Full article ">Figure 5
<p>Fraction of the remaining mass of injection-molded samples measured isothermally at 320 °C.</p>
Full article ">Figure 6
<p>Micrographs: pure materials (<b>a</b>) stereocomplex—ST and (<b>b</b>) TPS; sterecomplexes plus TPS; (<b>c</b>) ST + 15%TPS; (<b>d</b>) ST + 30%TPS; plus chain extender (<b>e</b>) ST + 15%TPS + 2%CE; and (<b>f</b>) ST + 30%TPS + 2%CE.</p>
Full article ">Figure 7
<p>Stress vs. strain for stereocomplex and TPS blends: “+2CE” labels samples with 2wt% chain extender added.</p>
Full article ">Figure 8
<p>Storage moduli of polylactide stereocomplex (ST) blends vs. temperature.</p>
Full article ">Figure 9
<p>Tan δ curves of polylactide stereocomplex (ST) blends vs. temperature.</p>
Full article ">Figure 10
<p>Heat resistance of injection-molded PLLA, TPS, and ST/TPS blends (<b>a</b>) as-fabricated; (<b>b</b>) after heating at 100 °C for 30 min.</p>
Full article ">Figure 11
<p>Deflection of the PLLA, TPS, ST blended with TPS, and chain extender under a 0.45 MPa load as the temperature increased.</p>
Full article ">
19 pages, 4280 KiB  
Article
High-Expansion Open-Cell Polylactide Foams Prepared by Microcellular Foaming Based on Stereocomplexation Mechanism with Outstanding Oil–Water Separation
by Dongsheng Li, Shuai Zhang, Zezhong Zhao, Zhenyun Miao, Guangcheng Zhang and Xuetao Shi
Polymers 2023, 15(9), 1984; https://doi.org/10.3390/polym15091984 - 22 Apr 2023
Cited by 5 | Viewed by 1717
Abstract
Biodegradable polylactic acid (PLA) foams with open-cell structures are good candidates for oil–water separation. However, the foaming of PLA with high-expansion and uniform cell morphology by the traditional supercritical carbon dioxide microcellular foaming method remains a big challenge due to its low melting [...] Read more.
Biodegradable polylactic acid (PLA) foams with open-cell structures are good candidates for oil–water separation. However, the foaming of PLA with high-expansion and uniform cell morphology by the traditional supercritical carbon dioxide microcellular foaming method remains a big challenge due to its low melting strength. Herein, a green facile strategy for the fabrication of open-cell fully biodegradable PLA-based foams is proposed by introducing the unique stereocomplexation mechanism between PLLA and synthesized star-shaped PDLA for the first time. A series of star-shaped PDLA with eight arms (8-s-PDLA) was synthesized with different molecular weights and added into the PLLA as modifiers. PLLA/8-s-PDLA foams with open-cells structure and high expansion ratios were fabricated by microcellular foaming with green supercritical carbon dioxide. In detail, the influences of induced 8-s-PDLA on the crystallization behavior, rheological properties, cell morphology and consequential oil–water separation performance of PLA-based foam were investigated systemically. The addition of 8-s-PDLA induced the formation of SC-PLA, enhancing crystallization by acting as nucleation sites and improving the melting strength through acting as physical cross-linking points. The further microcellular foaming of PLLA/8-s-PDLA resulted in open-cell foams of high porosity and high expansion ratios. With an optimized foaming condition, the PLLA/8-s-PDLA-13K foam exhibited an average cell size of about 61.7 μm and expansion ratio of 24. Furthermore, due to the high porosity of the interconnected open cells, the high-absorption performance of the carbon tetrachloride was up to 37 g/g. This work provides a facile green fabrication strategy for the development of environmentally friendly PLA foams with stable open-cell structures and high expansion ratios for oil–water separation. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H NMR spectra of synthesized star-shaped PDLA.</p>
Full article ">Figure 2
<p>(<b>a</b>) DSC heating curves of 8-s-PDLA with different molecular weights. (<b>b</b>) GPC curve for 8-s-PDLA-13KCharacterization of PLLA/8-s-PDLA Blends.</p>
Full article ">Figure 3
<p>(<b>a</b>) DSC second heating curves and (<b>b</b>) XRD patterns of PLLA-based blends with different contents of 8-s-PDLA-13K; (<b>c</b>) DSC second heating curves and (<b>d</b>) XRD patterns of PLLA-based blends with 8-s-PDLA-13K of different molecular weights.</p>
Full article ">Figure 4
<p>POM photographs of isothermal PLLA/8-s-PDLA-13K blends crystallized at 130 °C: (<b>a</b>) pure PLLA, (<b>b</b>) PLLA/8-s-PDLA-13K-3%, (<b>c</b>) PLLA/8-s-PDLA-13K-5%, and (<b>d</b>) PLLA/8-s-PDLA-13K-10%.</p>
Full article ">Figure 5
<p>Frequency dependence of storage modulus (<b>a1</b>), loss tangent angle (<b>a2</b>) and complex viscosity (<b>a3</b>) of PLLA with 8-s-PDLA of different molecular weights blended at 190 °C; Frequency dependence of storage modulus (<b>b1</b>), loss tangent angle (<b>b2</b>) and complex viscosity (<b>b3</b>) of PLA with 8-s-PDLA of different molecular weights blended at 200 °C; temperature dependence of storage modulus (<b>c1</b>), loss tangent angle (<b>c2</b>) and complex viscosity (<b>c3</b>) of PLA with 8-s-PDLA of different molecular weights at fixed frequency of 10 rad/s from 180 to 230 °C.</p>
Full article ">Figure 6
<p>SEM images of cell morphology of: (<b>a</b>) pure PLLA; (<b>b</b>) PLLA/8-s-PDLA-13K-5%; (<b>c</b>) PLLA/8-s-PDLA-15K-5%; and (<b>d</b>) PLLA/8-s-PDLA-39K-5%. The foaming condition was kept the same at 130 °C with a pressure of 20 MPa for 4 h, with following fast pressure release.</p>
Full article ">Figure 7
<p>Schematic illustration of microcellular foaming of (<b>a</b>) pure PLLA and (<b>b</b>) PLLA/8-s-PDLA considering the effect of SC-PLA on cell nucleation and cell growth.</p>
Full article ">Figure 8
<p>SEM images of cell morphology of: (<b>a</b>) PLLA/8-s-PDLA-13K-1%, (<b>b</b>) PLLA/8-s-PDLA-13K-3%, (<b>c</b>) PLLA/8-s-PDLA-13K-5% and (<b>d</b>) PLLA/8-s-PDLA-13K-1% + 39K-2% (the foaming condition was kept the same at 130 °C with a pressure of 20 MPa for 4 h, with following fast pressure release).</p>
Full article ">Figure 9
<p>SEM images of cell morphology for PLLA/8-s-PDLA-15K-5% prepared by different foaming temperatures: (<b>a</b>) 110 °C, (<b>b</b>) 120 °C, (c)125 °C, (<b>d</b>) 130 °C and (<b>e</b>) 135 °C when the pressure was fixed at 20 MPa for 2 h.</p>
Full article ">Figure 10
<p>SEM images of cell morphology of PLLA/8-s-PDLA-15K-5% prepared by different foaming pressures: (<b>a</b>) 16 MPa, (<b>b</b>) 20MPa and (<b>c</b>) 22 MPa at fixed temperature of 120 °C for 2 h.</p>
Full article ">Figure 11
<p>Adsorption capacity of the PLLA/8-s-PDLA foams for (<b>a</b>) CCl<sub>4</sub> and (<b>b</b>) silicone oil; related adsorption process for (<b>c</b>) CCl<sub>4</sub> and (<b>d</b>) silicone oil.</p>
Full article ">Figure 12
<p>Maximum adsorption capacity of series of PLLA/8-s-PDLA blends for CCl4 in blue column and silicone oil in orange column within 5 min with the corresponding SEM images of foam morphology.</p>
Full article ">
12 pages, 2448 KiB  
Article
A Biodegradable Stereo-Complexed Poly (Lactic Acid) Drinking Straw of High Heat Resistance and Performance
by Renzhi Li, Yangyang Feng, R. Hugh Gong and Constantinos Soutis
Materials 2023, 16(6), 2438; https://doi.org/10.3390/ma16062438 - 18 Mar 2023
Cited by 3 | Viewed by 3271
Abstract
Current biodegradable drinking straws suffer from poor heat resistance and rigidity when wet, causing user dissatisfaction. Here, a fully biodegradable straw formed by stereocomplexation of poly (lactic acid) (SC-PLA) is reported. Because of the unique strong interaction and high density of link chains [...] Read more.
Current biodegradable drinking straws suffer from poor heat resistance and rigidity when wet, causing user dissatisfaction. Here, a fully biodegradable straw formed by stereocomplexation of poly (lactic acid) (SC-PLA) is reported. Because of the unique strong interaction and high density of link chains between stereocomplex crystallites (over 70% crystallinity), SC-PLA straws outperform their counterparts on the market. This coupled with the advantages of simple processing (solution casting and annealing) and relatively low cost (~2.06 cents per straw) makes SC-PLA drinking straws a superior substitute for plastic ones. Commercially available PLLA straws lose almost 60% of their flexural strength when wet compared to less than 5% of the SC-PLA straws proposed in this study. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Photo of SC-PLA film and straw. (<b>b</b>) SEM image of unannealed SC-PLA film and (<b>c</b>) SEM image of annealed SC-PLA straw.</p>
Full article ">Figure 2
<p>Comparison of crystallisation and thermal behaviour of films and straws. (<b>a</b>) XRD patterns, (<b>b</b>) DSC diagram, and (<b>c</b>) FTIR spectrum.</p>
Full article ">Figure 3
<p>(<b>a</b>) Typical tensile stress-strain curves for all straws. (<b>b</b>) Typical wet tensile stress-strain curves for all straws immersed in hot water at 100 °C for 10 min. (<b>c</b>) Comparison of tensile strength of straws in different states.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>d</b>) State of three straws from each group after immersion in hot water at 100 °C for 10 min.</p>
Full article ">Figure 5
<p>(<b>a</b>) Typical flexural stress-strain curves for all straws. (<b>b</b>) Typical wet flexural stress-strain curves for all straws immersed in hot water at 100 °C for 10 min. (<b>c</b>) Comparison of flexural strength of straws in different states.</p>
Full article ">Figure 6
<p>Remaining weights of films and straws as a function of hydrolysis time.</p>
Full article ">
24 pages, 28969 KiB  
Article
Towards Self-Reinforced PLA Composites for Fused Filament Fabrication
by Neha Yadav, Tim Richter, Oliver Löschke, Bilen Emek Abali, Dietmar Auhl and Christina Völlmecke
Appl. Sci. 2023, 13(4), 2637; https://doi.org/10.3390/app13042637 - 18 Feb 2023
Cited by 5 | Viewed by 2368
Abstract
Aligned with the Sustainability Development Goals (SDGs), we present the complete methodology of preparing bio-based polymer filaments to be used in additive manufacturing, specifically by means of so-called Fused Filament Fabrication (FFF) in 3D printing. Filament production and 3D printing were both developed [...] Read more.
Aligned with the Sustainability Development Goals (SDGs), we present the complete methodology of preparing bio-based polymer filaments to be used in additive manufacturing, specifically by means of so-called Fused Filament Fabrication (FFF) in 3D printing. Filament production and 3D printing were both developed and optimised in this work. First, we focused on the steps of producing and optimising the extrusion process of unreinforced polylactic acid (PLA) composite filaments. Second, we studied the resulting material properties by discussing the selection of a specimen geometry and the international standards adequate for FFF 3D printing. Moreover, we investigated the process parameters in order to achieve reliable structures. Based on the reinforcement material (stereocomplex fibres (Sc-PLA fibre) and bi-component fibres (bi-co PLA fibre), base-matrices were selected for producing un-reinforced filaments. In this way, we present the complete preparation approach by identifying problems and pitfalls for fostering studies of bio-based polymer filaments. Full article
(This article belongs to the Special Issue Advanced Materials in 3D Printing)
Show Figures

Figure 1

Figure 1
<p>Lactic acid: derived from fermentation of starch, the monomer of the polymer polylactic acid (PLA).</p>
Full article ">Figure 2
<p>L−lactic acid (<b>left</b>) and D−lactic acid (<b>right</b>), the two enantiomeric forms of the monomer lactic acid.</p>
Full article ">Figure 3
<p>Visualisation of the underlying manuscript’s targeted research areas for the production of the PLA composites used in additive manufacturing.</p>
Full article ">Figure 4
<p>(<b>a</b>) visualisation showing core PLA and matrix PLA for the Sc-PLA PLA composite, (<b>b</b>) visualisation showing core bi-co fibre in the PLA matrix for the second self-reinforced composite, and (<b>c</b>) cross section of the bi-co fibre showing the inner core and the outer sheath material construction. The mechanical properties are mentioned in <a href="#applsci-13-02637-t002" class="html-table">Table 2</a>, <a href="#applsci-13-02637-t003" class="html-table">Table 3</a> and <a href="#applsci-13-02637-t005" class="html-table">Table 5</a>.</p>
Full article ">Figure 5
<p>Extrusion set-up assembly constituting the production of (un)reinforced PLA filaments. (1) Hopper, (2) nozzle for the potential reinforced filament production, (3) processing screw with 3 sections (feeding, compression and metering), (4) water bath, (5) conveyor belt, (6) laser diameter controller, (7) filament winding station.</p>
Full article ">Figure 6
<p>Schematic diagram of the filament production: Extrusion set-up assembly constituting the production of (un)reinforced PLA filaments. (1) Hopper, (2) nozzle for the potential reinforced filament production, (3) processing screw with 3 sections (feeding, compression and metering), (4) water bath, (5) conveyor belt, (6) laser diameter controller, (7) filament winding station.</p>
Full article ">Figure 7
<p>Placement of the nozzle in the extrusion processing machine for the production of future continuous fibre-reinforced polymer composites.</p>
Full article ">Figure 8
<p>Microscopic images of base-matrix material A filament, PLA L130. (<b>Top</b>): challenge such as voids and unmelted polymer during filament production. (<b>Bottom</b>): good quality filament.</p>
Full article ">Figure 9
<p>Microscopic images of base-matrix material B filament, PLA 6302D. (<b>Top</b>): challenge such as voids and unmelted polymer during filament production. (<b>Bottom</b>): good quality filament.</p>
Full article ">Figure 10
<p>(<b>Top</b>): Microscopic image of unmelted polymer for base matrix materials A and B. (<b>Bottom</b>): Photos of unmelted granules while filament processing.</p>
Full article ">Figure 11
<p>Geometry of the tensile specimen in accordance with the ASTM D3039 used for testing unreinforced specimens. Commonly used in technical drawings, A and B are the reference surfaces for the parallel tolerances as seen in the ASTM D3039.</p>
Full article ">Figure 12
<p>Although the filament diameter is measured and regulated during filament production between fixed values of 1.65 mm to 1.85 mm via cross-axis laser micrometer, see <a href="#applsci-13-02637-f005" class="html-fig">Figure 5</a> and <a href="#applsci-13-02637-f006" class="html-fig">Figure 6</a>, it is additionally verified and re-structured. The example shows pure PLA material A (Luminy L130), which is pulled through the borehole number 19 with a 1.9 mm diameter for high quality 3D printing filaments with a final nominal value of 1.75 mm. This quality verification step becomes especially important for self(reinforced) PLA, as the inner-core may stick out or surface distortions can occur due to core/shell/sheath structure.</p>
Full article ">Figure 13
<p>Three dimensional (3D) printed device to keep the filament coming from water bath on to the conveyor belt. (<b>a</b>) front-view of the 3D printed structure to keep the filament from slipping off of the conveyor belt, (<b>b</b>) top-view of the 3D printed structure.</p>
Full article ">Figure 14
<p>The diagonal travel path of the print head across a 90° orientated specimen, displayed in light blue.</p>
Full article ">Figure 15
<p>Optimised specimen placement for a 90° orientated specimen, which can also be applied to 0° specimens.</p>
Full article ">Figure 16
<p>Optimised travel moves on a −45°/45° stacked specimen with the implementation of auxiliary models.</p>
Full article ">Figure 17
<p>The 3D printed 0°, 90° and −45°/45° orientated specimens of the type A material.</p>
Full article ">Figure 18
<p>The 3D printed 0°, 90° and −45°/45° orientated specimens of the type B material.</p>
Full article ">
20 pages, 5206 KiB  
Article
Scalable Continuous Manufacturing Process of Stereocomplex PLA by Twin-Screw Extrusion
by Mohammed Alhaj and Ramani Narayan
Polymers 2023, 15(4), 922; https://doi.org/10.3390/polym15040922 - 12 Feb 2023
Cited by 3 | Viewed by 2958
Abstract
A scalable continuous manufacturing method to produce stereocomplex PLA was developed and optimized by melt-blending a 1:1 blend of high molecular weight poly(L-lactide) (PLLA) and high molecular weight poly(D-lactide) (PDLA) in a co-rotating twin-screw extruder. Thermal characteristics of stereocomplex formation were characterized via [...] Read more.
A scalable continuous manufacturing method to produce stereocomplex PLA was developed and optimized by melt-blending a 1:1 blend of high molecular weight poly(L-lactide) (PLLA) and high molecular weight poly(D-lactide) (PDLA) in a co-rotating twin-screw extruder. Thermal characteristics of stereocomplex formation were characterized via DSC to identify the optimal temperature profile and time for processing stereocomplex PLA. At the proper temperature window, high stereocomplex formation is achieved as the twin-screw extruder allows for alignment of the chains; this is due to stretching of the polymer chains in the extruder. The extruder processing conditions were optimized and used to produce >95% of stereocomplex PLA conversion (melting peak temperature Tpm = 240 °C). ATR-FTIR depicts the formation of stereocomplex crystallites based on the absorption band at 908 cm−1 (β helix). The only peaks observed for stereocomplex PLA’s WAXD profile were at 2θ values of 12, 21, and 24°, verifying >99% of stereocomplex formation. The total crystallinity of stereocomplex PLA ranges from 56 to 64%. A significant improvement in the tensile behavior was observed in comparison to the homopolymers, resulting in a polymer of high strength and toughness. These results lead us to propose stereocomplex PLA as a potential additive/fiber that can reinforce the material properties of neat PLA. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Enantiomeric PLA homopolymers blend to form stereocomplex PLA.</p>
Full article ">Figure 2
<p>The first DSC thermogram of pellets comprising 50% stereocomplex crystallites/50% PLA homocrystallites based on the melting enthalpy calculations.</p>
Full article ">Figure 3
<p>Processing conditions for the first trial run of stereocomplex PLA in a co-rotating twin-screw extruder type ZSE 27 HP–PH from Leistritz.</p>
Full article ">Figure 4
<p>The DSC thermograms on stereocomplex PLA (<b>a</b>) First heating scan—stereocomplex crystallite formation (<b>b</b>) Second heating scan—thermal dissociation of stereocomplex crystallites to homocrystallites.</p>
Full article ">Figure 5
<p>Processing conditions for the new trial run of stereocomplex PLA in a co-rotating twin-screw extruder type ZSE 27 HP–PH from Leistritz.</p>
Full article ">Figure 6
<p>The DSC thermograms displaying stereocomplex formation after the sample was held isothermally for 1–3 min at temperatures of (<b>a</b>) 180 °C, (<b>b</b>) 190 °C, (<b>c</b>) 200 °C, (<b>d</b>) 210 °C, (<b>e</b>) 220 °C, (<b>f</b>) 230 °C.</p>
Full article ">Figure 7
<p>(<b>a</b>) Isothermal stereocomplexation kinetics. (<b>b</b>) Effect of temperature on stereocomplex formation.</p>
Full article ">Figure 8
<p>(<b>a</b>) FTIR spectra of the PLLA (black line) and stereocomplex PLA (red line). (<b>b</b>) FTIR spectra displaying characteristic absorption bands. The spectra displayed involves the stereocomplex PLA collection time at 20 min.</p>
Full article ">Figure 9
<p>WAXD profiles of (<b>a</b>) pure PLLA and (<b>b</b>) stereocomplex PLA (50/50 L/D).</p>
Full article ">Figure 10
<p>The DSC thermograms of SC PLA at sample collection times of (<b>a</b>) 5 min, (<b>b</b>) 10 min, (<b>c</b>) 20 min, (<b>d</b>) 40 min.</p>
Full article ">Figure 11
<p>The tensile stress–strain curves of injection molded samples of neat PLLA (L175) and stereocomplex PLA.</p>
Full article ">Figure 12
<p>Comparison of the tensile behavior of extruded stereocomplex PLA vs. PLLA and the literature results. (<b>a</b>) Yield strength, (<b>b</b>) ultimate tensile strength, (<b>c</b>) elastic modulus, (<b>d</b>) elongation at break.</p>
Full article ">Figure 13
<p>Thermal degradation of (<b>a</b>) stereocomplex PLA and (<b>b</b>) PLLA (L175).</p>
Full article ">
Back to TopTop