Open Access Article
Theodosis
Giousis†
ab,
Zoi
Terzopoulou
*c,
Maria-Eirini
Grigora
d,
Dimitrios
Moschovas
b,
Stamatia
Spyrou
e,
Renia
Fotiadou‡
e,
Haralambos
Stamatis
e,
Apostolos
Avgeropoulos
b,
Dimitrios
Tzetzis
d,
Dimitrios P.
Gournis
bf,
Dimitrios N.
Bikiaris
g and
Petra
Rudolf
*a
aZernike Institute for Advanced Materials, University of Groningen, Nijenborgh 3, 9747 AG Groningen, The Netherlands. E-mail: p.rudolf@rug.nl
bDepartment of Materials Science & Engineering, University of Ioannina, 45110 Ioannina, Greece
cLaboratory of Industrial Chemistry, Department of Chemistry, University of Ioannina, 45110 Ioannina, Greece. E-mail: terzoi@uoi.gr
dDigital Manufacturing and Materials Characterization Laboratory, School of Science and Technology, International Hellenic University, 14 km Thessaloniki, 57001 N. Moudania, Greece
eBiotechnology Laboratory, Department of Biological Applications and Technologies, University of Ioannina, 45110 Ioannina, Greece
fSchool of Chemical and Environmental Engineering, Technical University of Crete, 73100 Chania, Crete, Greece
gLaboratory of Polymers and Colours Chemistry and Technology, Department of Chemistry, Aristotle University of Thessaloniki, GR-54124 Thessaloniki, Greece
First published on 13th October 2025
Poly(lactic acid) (PLA) is a widely used biobased polymer, but its slow crystallization, brittleness, and limited functional properties restrict broader applications. In this study, we report the first incorporation of germanane (GeH) into PLA via solution mixing to produce nanocomposites. Adequate dispersion was achieved at low GeH loadings (0.5–3.0 wt%), while higher concentrations (5.0 wt%) led to aggregation. The addition of small amounts of GeH significantly accelerated PLA crystallization and enhanced local mechanical properties, although thermal stability was slightly reduced. Notably, the nanocomposites exhibited antioxidative and antibacterial activities, arising from the intrinsic properties of GeH. These findings highlight that very low GeH loadings are sufficient to enhance both structural and functional performance. The combination of improved crystallization, mechanical behavior, and bioactive properties positions PLA/GeH nanocomposites as promising candidates for applications in bioactive packaging and biomedical materials.
Following the success of graphene, attention has turned to other 2D monoelemental Xenes, such as phosphorene, antimonene, silicene, and germanene. Germanane (GeH), which is used in this study, is the hydrogenated counterpart of germanene and possesses several distinctive advantages over conventional 2D fillers. Its hydrogen-terminated surface enables facile functionalization for strong interfacial adhesion with PLA without harsh treatments. Unlike defect-rich graphene oxide or metallic MXenes, GeH exhibits a direct bandgap (∼1.6 eV) and high carrier mobility, which allow visible-light activity, reactive oxygen species generation, and simultaneous antioxidant and antibacterial effects. Its nonmetallic composition reduces cytotoxicity risks, and it can be dispersed under mild conditions while preserving PLA's transparency and minimizing embrittlement at low loadings.13,14
Germanane was first synthesized via topochemical reaction of CaGe2 with HCl at low temperature.15,16 Earlier methods were time-consuming, and produced material with limited thermal stability and purity. Our previously developed synthetic protocol yields highly pure, thermally stable GeH in minutes,17 which is critical for high-performance PLA nanocomposites.
Germanane has been reported to possess significant antibacterial activity against both Gram-negative and Gram-positive bacterial strains.18 Despite the intrinsic mechanical deficiencies of PLA, its potential in food packaging applications increases significantly when stabilized or enriched with antioxidants.19–21 Other polymer-based nanocomposites have been reported as efficient carriers for antioxidants22,23 or for skin and bone tissue engineering.24,25 Among these properties, antimicrobial activity is particularly important for biomedical applications. To date, antimicrobial polymeric nanocomposites reinforced with graphene derivatives26 or copper27 have been reported; however, studies on polymer-based nanocomposites incorporating functionalized germanene remain scarce. Feng et al.28 prepared hydrogel-based nanocomposites with drug-loaded PEGylated Ge nanosheets for cancer treatment, demonstrating efficient postoperative wound coating and excellent theranostic properties.
PLA/GeH nanocomposites are thus promising for applications requiring combined mechanical reinforcement and bioactivity, such as bioactive packaging and biomedical devices. This study addresses this knowledge gap by evaluating the structural, thermal, crystallization, mechanical, and bioactive properties of PLA/GeH nanocomposites. To explore GeH as a nanofiller for PLA, nanocomposites with different GeH content (0.5, 1.0, 2.0, 3.0, and 5.0 wt%) were prepared by solution casting. The nanocomposites are hereafter referred to as PLA/GeH 0.5, PLA/GeH 1.0, PLA/GeH 2.0, PLA/GeH 3.0, and PLA/GeH 5.0. The effects of the GeH nanosheets on the physicochemical, mechanical, and crystallization properties of PLA were investigated using Fourier transform infrared (FTIR) spectroscopy, X-ray diffraction (XRD), thermogravimetric analysis (TGA), differential scanning calorimetry (DSC), energy-dispersive X-Ray spectroscopy (EDS), and nanoindentation. Transmission electron microscopy (TEM) and scanning electron microscopy (SEM) were employed to examine the morphology and structure of the nanocomposites. Finally, the biological properties of the prepared nanocomposites were evaluated, including their antioxidant activity via ABTS and DPPH radical scavenging assays and their antimicrobial activity against Escherichia coli (BL21(DE3)) and Corynebacterium glutamicum (ATCC 21253).
700 g mol−1 and Mw = 180
300 g mol−1 (SEC), specific gravity 1.24 g per cc, melt flow rate (MFR) 6 g/10 min at 210 °C) from Natureworks was kindly donated by Plastika Kritis S.A. LB Broth Lennox was purchased from NEOGEN and sodium chloride from Riedel de Haen. All other chemicals were of reagent grade and purchased from Sigma-Aldrich.
![]() | ||
| Scheme 1 Schematic representation of the experimental procedures followed for the fabrication of the PLA/GeH nanocomposites and the evaluation of the bactericidal properties. | ||
Atomic force microscopy images of GeH were obtained in AC mode using an Asylum Cypher-S instrument, Asylum Research, with HQ-300 Au-coated cantilevers with a tip radius of 10 nm, spring constant of 40 N m−1 and operating frequency of 300 kHz. Exfoliated GeH nanosheets were drop casted onto silicon wafers (P/Bor, single side polished, purchased from Si-Mat) from ethanol dispersions.
The FTIR spectrum of GeH was acquired with a Shimadzu FTIR 8400 spectrometer equipped with a deuterated triglycine sulfate (DTGS) detector in the range of 400–4000 cm−1, averaging 32 scans collected with 2 cm−1 resolution. The sample was in the form of a KBr pellet containing ca. 2 wt% of GeH. FTIR spectra of PLA/GeH thin films, prepared by spin-coating onto SiO2 substrates, were collected using a PerkinElmer SPECTRUM 1000 FTIR spectrometer. The resolution was 2 cm−1, and the number of co-added scans in each spectrum was 16; the spectra presented were baseline-corrected and converted to absorbance mode. Raman spectra of GeH were obtained using a Labram Horiba HR spectrometer with a laser excitation wavelength of 514 nm; the laser power of 1.5 mW was focused onto a 2 μm spot.
The XRD pattern of GeH was collected on a D8 Advance Bruker diffractometer with Cu Kα radiation (40 kV, 40 mA) and a secondary-beam graphite monochromator. The pattern was recorded in the 2θ range of 10–65°, in steps of 0.02°, and a counting time of 2 s per step. XRD measurements of the nanocomposites were performed over the 2θ range of 5 to 60°, in steps of 0.05°, scanning speed 1° min−1, using a MiniFlex II XRD system from Rigaku Co. with Cu Kα radiation (λ = 0.154 nm). Transmission electron microscopy (TEM) was performed using a JEOL JEM HR-2100 instrument, operated at 120 keV. The drop-casted PLA-GeH5.0 film was cryo-ultramicrotomed using a a Leica EM UC7, producing (∼40 nm thin sections) that were picked up on 600-mesh copper TEM grids for immediate observation.
Themogravimetric analysis (TGA) was carried out with a Setaram Setsys TG-DTA 16/18 instrument. Samples (8 ± 0.2 mg) were placed in alumina crucibles; a blank measurement was performed and subsequently subtracted from the experimental curves of GeH and of the PLA/GeH nanocomposites to correct for buoyancy effect. The nanocomposite samples were heated from ambient temperature up to 600 °C with a heating rate of 20 °C min−1, while GeH samples were heated up to 1000 °C, under a 50 mL min−1 N2 flow. Both sample temperature and sample weight were continuously recorded.
Differential scanning calorimetry (DSC) studies were performed using a PerkinElmer (Shelton, Connecticut, USA) Pyris Diamond DSC calorimeter under a nitrogen gas flow of 50 mL min−1. The instrument was calibrated with indium for the accurate determination of heat flow and temperature. A sample mass of 5.00 mg was used for all tests; the sample and reference pans were of identical mass within ±0.01 mg. The degree of crystallinity (Xc) was calculated with eqn (1):
![]() | (1) |
Isothermal crystallization from the melt experiments were performed at temperatures from 97.5 to 107.5 °C. PLA was first melted at T = Tm + 40 °C for 5 min to erase all thermal history and then cooled at a rate of 200 °C min−1 to the desired crystallization temperature. After holding isothermally until crystallization was complete, a heating step with a rate of 20 °C min−1 to T = Tm + 40 °C followed.
The mechanical performance of the PLA/GeH nanocomposites was investigated via nanoindentation testing. The samples were indented with a 100 nm radius triangular pyramidal tip (Berkovich – type indenter) mounted on a dynamic ultra-micro-hardness tester (DUH-211; Shimadzu Co., Kyoto, Japan). These tests precisely measure local variations of elastic modulus and hardness;29–31 the load of the indenter was recorded as a function of the indentation depth. The measurements were carried out in five different points on the surface of each film, with a holding time of 3.0 s at both load and unload. Nanoindentation tests were performed in the load-control mode at a peak load of 50 mN. During the creep time the maximum indentation load was applied to the indenter; the change in indentation depth (displacement) was monitored as a function of time and subsequently the indenter was unloaded until zero load.
The antioxidant activity of the PLA surfaces was evaluated based on the DPPH and ABTS radical scavenging assays,32–34 PLA and PLA/GeH nanocomposite films of approximately 0.25 cm2 were immersed in ethanolic DPPH˙ solution (0.03 mM) and left at RT. Measurements were conducted using spectrophotometry at 517 nm, with ethanol as the blank. For the ABTS assay, PLA and PLA/GeH nanocomposite films (0.25 cm2) were immersed in aqueous ABTS˙+ solution adjusted to a final absorbance of 0.7 ± 0.05, and incubated at room temperature for different time intervals (30, 60, 120, 240 min). The decrease in absorbance of the ABTS˙+ solution was measured at 734 nm using a UV/vis microplate reader (Multiskan SkyHigh, Thermo Fisher Scientific, Cleveland, OH, USA). All experiments were conducted in triplicate and results are expressed as mean ± standard deviation. The radical scavenging activity (RSA) was calculated according to the following equation (eqn (2)):
![]() | (2) |
Statistical analysis was performed by analysis of variance (ANOVA). Multiple comparisons of means (Dunnett's and Tukey's tests) were conducted to identify significant differences, which were considered at p < 0.05, using IBM SPSS Statistics version 21 (SPSS Inc., Chicago, IL, USA).
The antibacterial activity of the PLA/GeH nanocomposite films was tested against a Gram-negative strain (Escherichia coli-E. coli) and a Gram-positive one (Corynebacterium glutamicum-C. glutamicum), as previously reported.33 In brief, a fresh bacterial inoculum was prepared by incubating bacterial cells overnight in sterile Lysogeny Broth (LB) at 37 °C with shaking. An exponential-phase bacterial population corresponding to ≈107 CFU mL−1 was then prepared in 0.9% NaCl solution (w/v), and 100 μL was added to Eppendorf tubes containg 0.25 cm2 of each film, followed by incubation for 18 h in a cold chamber. Control samples consisted of 100 μL of bacteria incubated without a nanocomposite film. After 18 h, 25 μL from each sample were transferred to a 96-well Elisa microplate containing 225 μL of LB medium. The microplate was incubated at 37 °C for 8 h with continuous stirring, and bacterial growth was determined by measuring the Optical Density (O.D.) at 600 nm at 1 h intervals. The % lethal effect of each sample was calculated according to the following equation (eqn (3)):
![]() | (3) |
To further elucidate the structure and the chemical composition of the pristine GeH, Raman and FTIR spectra were recorded. Raman spectrum of GeH (Fig. 1(b)) is dominated by an intense peak at 287 cm−1, which corresponds to the E2g in-plane vibration mode of the Ge atoms in the honeycomb lattice of GeH.16,17 In the FTIR spectrum of germanane (Fig. 1(c)) the strong peak at 2000 cm−1 that derives from the Ge–H stretching vibration, as well as the signature of the Ge–H wagging modes at 480 and 584 cm−1 confirm that the germanium atoms are hydrogenated.16,17 In addition, the weak peaks at 770 and 830 cm−1 stem from H–Ge–H bending modes of neighboring H-terminated germanium atoms at the edges of the crystalline layers and/or next to Ge vacancies in the germanane lattice.16,17 Scanning electron and atomic force microscopy images revealed the layered structure of the GeH crystallites. Both panels, (d and e) in Fig. 1 present the SEM observations of different GeH flakes. Fig. 1(d) shows two distinctive GeH crystallites (top view), while Fig. 1(e) shows the side view of another GeH crystal, highlighting the layered nature and the accordion-like structure of the sample. Fig. 1(f) present the AFM image of different exfoliated individual GeH nanosheets and their height profile analysis is presented in Fig. 1(g). The thickness of a single layer germanane was found to be 0.9 nm ± 0.1 nm, in agreement with previous studies.17
![]() | ||
| Fig. 2 SEM images of the surface (left) and the cross section (right) of bulk samples of (a) and (ai) neat PLA; (b) and (bi) PLA/GeH0.5; (c) and (ci) PLA/GeH2.0 and (d) and (di) PLA/GeH5.0. | ||
For the cross section of the PLA/GeH 5.0 nanocomposite, shown in Fig. 3(a), energy-dispersive X-ray spectroscopy (EDS) data were acquired, and are presented in Fig. 3(b). The elemental mapping shown in Fig. 3(ai)–(aiv) indicates that the Ge signal clearly overlaps everywhere with the carbon and oxygen signals coming from (C3H4O2)n. This observation confirms both the successful incorporation of GeH and the uniform distribution of the nanofiller in the PLA matrix. In the EDS elemental mapping of Ge (Fig. 3(aiii)) one also notices a few high intensity spots, which point to the presence of not perfectly exfoliated flakes of GeH. Fig. 3(c) and (d) present the TEM images of the nanocomposite PLA/GeH 5.0. The images show homogeneously dispersed, exfoliated GeH incorporated in the PLA polymer matrix (parts identified with white ovals in Fig. 3(d)). However, careful observation of the microtomed sections in TEM revealed again not completely exfoliated GeH flakes as identified in Fig. 3(d) with a red oval.
To determine whether any specific bonds are formed between PLA and the GeH nanosheets in the composite, FTIR spectra were recorded (Fig. S1(a)). The main bands in the spectrum of PLA are observed at 3507 cm−1 (O–H stretching), 2995 cm−1 (–CH3 asymmetric stretching), 2945 cm−1 (–CH symmetric stretching), 1757 cm−1 (C
O stretching), 1453 cm−1 (–CH3 bending), 1382 cm−1 (–CH3 scissor mode), 1184 cm−1 (C–O stretching), and 1090 cm−1 (C–CH3 stretching).35 The positions of the FTIR bands remain unchanged after the incorporation of GeH, indicating that no new bonds were formed between the polymer and the nanofiller.
The X-Ray diffraction patterns of bulk films of both PLA and its nanocomposites, shown in Fig. S1(b), reveal an amorphous structure in all cases, with the typical broad halo at 15–20°. This result is expected given the high molecular weight of the PLA used and its very slow crystallization rate. Notably, a small peak appears at 27.7° upon addition of GeH nanosheets, and its intensity increases with higher GeH content. Since GeH nanosheets are highly crystalline, the reflections at ∼17° and 27.7° correspond to the (002) and (001) crystallographic planes, respectively.16,17 The (002) peak is barely visible in the nanocomposites with low nanofiller content, likely because too few non-exfoliated GeH flakes were present to give rise to a measurable signal, but it is clearly seen in the sample with 5 wt% nanofiller. This suggests that the degree of exfoliation of the nanofiller seems is influenced by its concentration in the nanocomposite. The presence of non-exfoliated GeH flakes in the PLA matrix agrees with the TEM and the SEM findings presented in Fig. 3.
![]() | ||
| Fig. 4 Effect of GeH content on the melting point, Tm, the glass transition temperature, Tg, and the crystallinity Xc of PLA. | ||
The resulting mass (%)–temperature curves and differential thermogravimetric analysis (DTG) graphs from TGA are depicted in Fig. S3. The characteristic peak temperature of the DTG curve, where decomposition occurs at the fastest rate (Tp), together with the percentage of char residue, are presented in Table 1. A first mass loss of 2–5%, depending on GeH content, is barely perceptible in neat PLA but clearly visible in the nanocomposites, occurring at Tp ≈ 150 °C. This loss corresponds to the release of residual solvent from solution casting. The main degradation of PLA takes place in a single step between 350 and 400 °C. The nanocomposites follow a similar trend but with a slightly reduced Tp. This small decrease in thermal stability may result from the lack of strong interactions between the PLA matrix and GeH nanosheets.37,38 The solid residue increases with higher GeH content, as expected due to the inorganic nature of the latter.
| Sample | T p (°C) | Residue at 600 °C (wt%) |
|---|---|---|
| PLA | 387.6 | 0.88 |
| PLA/GeH 0.5 | 383.1 | 1.06 |
| PLA/GeH 1.0 | 375.2 | 1.56 |
| PLA/GeH 2.0 | 380.8 | 1.80 |
| PLA/GeH 3.0 | 371.7 | 2.55 |
| PLA/GeH 5.0 | 375.6 | 3.75 |
The crystallinity of the polymer in the nanocomposite not only affects the optical properties but also directly influences the processing when manufacturing different items such as films and bottles. Crystallization of PLA is essential before drying the pellets to avoid clumping, or after molding into components to enhance mechanical performance. Increasing the crystallinity not only improves barrier properties but also stabilizes the shape of PLA-based items.39 PLA of medium to high molecular weight crystallizes very slowly;39–43 according to the manufacturer, the optimum conditions for the crystallization of PLA 2003D, used in this study, are 88–99 °C for 10–20 min. Nanoparticles are known for their ability to accelerate polymer crystallization by acting as heterogeneous nucleation agents, thereby increasing the nucleation density of PLA.40,44–47
The isothermal melt crystallization kinetics of the PLA nanocomposites were studied at crystallization temperatures between 97.5 and 107.5 °C. The resulting differential scanning calorimetry curves recorded during the isothermal step and subsequent melting are presented in Fig. S4 and S5, respectively. The evolution of the relative crystallinity, X(t), of neat PLA and its composites as a function of time at different crystallization temperatures is displayed in Fig. S6. Fig. 5(a), presents the inverse of the crystallization half time, t1/2, defined as the time required for a sample to reach 50% of its total crystallinity, while Fig. 5(b) shows the final Xc for neat PLA and the nanocomposites. The t1/2 of PLA ranged from 26 to 34 min (Table S1), in good agreement with previously reported values for PLA with the same D-isomer content and similar molecular weight.43,48 The spherulite growth rate of PLA is known to reach a minimum when the molecular weight is ≥100
000 g mol−1 because of restricted chain mobility.40 For neat PLA, both the time to peak and t1/2 increased slightly with temperature, although the effect was not significantly. By contrast, the nanocomposites especially those containing 0.5–3.0 wt%-consistently exhibited lower t1/2 values across the temperature range. Adding 0.5, 1.0 and 3.0 wt% GeH reduced the t1/2 of PLA by about 5 min in the temperature range 97.5–102.5 °C, probably because the GeH nanosheets act as nucleation agent. Conversely, t1/2 was not greatly affected after the incorporation of 5.0 wt% GeH, likely due to the presence of aggregates. Consequently, the nanocomposites with GeH content between 0.5 and 3.0 wt% showed an increased Xc by approximately 2–4% after melt crystallization in the 97.5–102.5 °C range (Fig. 5(b)), whereas the effect was less pronounced at higher temperatures. The decrease of crystallinity in the presence of 5.0 wt% GeH can be a consequence of confinement, supported by Xc dropping to ∼21–23%.49 This trend suggests that at higher filler loadings, the GeH flakes may hinder chain mobility and hinder crystalline growth.
The isothermal crystallization kinetics were analyzed with the Avrami method (detailed in the SI). The effect of the melt crystallization temperature and of the presence of the fillers on the Avrami exponent n and the growth function k is shown in (Fig. 5(c) and (d)). Herein, the Avrami exponent n ranged between 1.90 and 2.35 (Fig. 5(c)) and slightly increased with crystallization temperature. The presence of 0.5–3.0 wt% GeH generally led to higher n values, implying a shift toward more complex (likely 3D) crystal growth with simultaneous heterogeneous nucleation.43 For PLA/GeH 5.0, a lower n value was observed, consistent with the longer t1/2 times and lower crystallinity. This suggests that high filler loadings might hinder crystallization due to confinement and reduced chain mobility.49–57
The crystallization rate constant k (Fig. 5(c)) also increased with the addition of 0.5–3.0 wt% GeH, indicating faster crystallization, again confirming the role of GeH as a nucleation agent.45,58–61 Both k and the inverse t1/2 (Fig. 5(a) and (c)) followed a similar trend, generally decreasing at higher temperature. The highest crystallization rate was observed at 100 °C in the PLA/GeH 3.0 wt% sample. The layered structure of GeH provides a high specific surface area that facilitates heterogeneous nucleation by offering planar surfaces on which PLA chains can adsorb, align and crystallize. This geometry-driven nucleation is more effective than simply serving as a physical barrier, as it promotes oriented crystal growth and accelerates spherulite formation without significantly affecting Tg or Tm. At higher loadings (≥5 wt%), aggregation reduces the effective surface area, diminishing nucleation efficiency and slightly limiting the crystallinity enhancement.
Although no clear evidence of hydrogen bonding between PLA ester –C
O groups and GeH nanoadditives was observed in the FTIR spectra, the thermal analyses suggest that weak noncovalent interactions occur. These interactions appear sufficient to promote nucleation, leading to faster crystallization and slightly higher crystallinity at low GeH loadings (0.5–3 wt%), while having minimal impact on Tg, Tm, and overall thermal stability. Despite their low strength, these interactions help maintain the stability of the composites and prevent significant migration of the nanosheets under normal conditions. Consequently, PLA/GeH nanocomposites could prove suitable for applications requiring moderate thermal and mechanical performance, such as bioactive packaging and biomedical devices.
The elastic modulus (Fig. 6(b)) for neat PLA was 2484.8 MPa. The addition of 0.5 wt% GeH increased the modulus to 3815.3 MPa, which corresponds to 53% enhancement compared to neat PLA. Adding 1.0 and 2.0 wt% GeH to PLA also increased the elastic modulus by 20% and 15%, respectively. In contrast, PLA with 3.0 wt% and 5.0 wt% GeH exhibited elastic moduli that were 5% and 8% lower than that of neat PLA. This decrease can be attributed to agglomeration of the nanofiller within the PLA matrix.
Fig. 6(c) compares the load-depth curves of neat PLA and PLA/GeH nanocomposites as obtained from the nanoindentation tests. As expected from the hardness results, the indentation depths were highest for neat PLA (3.53 to 3.90 μm) and lowestfor PLA/GeH 0.5 (3.12 to 3.28 μm), while the other PLA/GeH samples showed intermediate values. Fig. 6(d) illustrates the creep displacement at the peak force of 50 mN as a function of time. The creep displacement is the difference between the indentation depth at the moment when the peak load of 50 mN is reached and the indentation depth at the end of holding time under constant load. The variation of the creep displacement with holding time confirms that the addition of GeH has a positive impact on the hardness since the creep displacement was significantly reduced compared to neat PLA. It seems that the GeH nanosheets act as blocking sites, hindering the movement of polymer chains subjected to the deformation field. This effect becomes particularly evident at higher GeH concentrations, where the nanoplatelets restrict the viscous flow of PLA. Moreover, the significant reduction in creep displacement even at the highiest GeH concentration demonstrates efficient load transfer between the PLA matrix and the nanomaterial, which is at the origin of the good creep resistance of the nanocomposites. While the solvent-induced dense spherulites observed by SEM may contribute to mechanical behavior, the systematic trends in hardness, elastic modulus, and creep resistance showed correlation with GeH content and dispersion.
The antibacterial activity of the PLA/GeH nanocomposite films was tested against E. coli and C. glutamicum, representing Gram-negative and Gram-positive strains, respectively. As shown in Fig. 7(c), neat PLA exhibited only slight antibacterial activity against both strains. Germanane is known to possess antibacterial properties against both Gram-positive and Gram-negative bacteria.18 Consequently, the PLA/GeH nanocomposites were evaluated for GeH-loading-dependent antibacterial activity. The results show that the antibacterial effect increased with GeH content, with the highest activity observed at 5 wt% GeH (71.3% against E. coli and 64.8% against C. glutamicum). Interestingly, the nanocomposites exhibited slightly stronger antibacterial activity against the Gram-negative strain than the Gram-positive one. The observed antibacterial action is likely due to cell membrane damage caused by germanane, as previously reported.18 The “sharp” edges of the nanosheets at the surface of the nanocomposite cut through the adsorbed bacterium's cell membrane, causing the intracellular matrix to leak, which eventually leads to the bacterium's death.
1278::AID-ADMA1278
3.0.CO;2-Y.
nm monolayer germanane transistors: a first-principle quantum transport simulation, J. Appl. Phys., 2024, 135(13), 134303, DOI:10.1063/5.0192389.Footnotes |
| † Present address: Department of Chemistry, University of Ioannina, GR 45110, Ioannina, Greece. |
| ‡ Present address: Department of Chemistry, University of Crete, GR-70013, Heraklion, Crete, Greece. |
| This journal is © The Royal Society of Chemistry 2025 |