Bioactive potential of marine Aspergillus niger AMG31: Metabolite profiling and green synthesis of copper/zinc oxide nanocomposites – An insight into biomedical applications
-
Rana Hussein Naser
Abstract
This investigation explored marine fungi from Red Sea sediments, focusing on Aspergillus niger AMG31. Chemical profiling of the fungal extract by high-performance liquid chromatography and gas chromatography–mass spectrometry revealed diverse bioactive compounds, with hesperetin (80,471.56 μg·g−1) and rosmarinic acid (8,396.08 μg·g−1) predominating. Additionally, the extract contained substantial phenolics (55.517 mg·g−1), flavonoids (28.757 mg·g−1), and tannins (18.650 mg·g−1). The fungal extract facilitated green synthesis of copper-zinc oxide nanocomposites (CZ nanocomposites), which were thoroughly characterized using fourier transform infrared spectroscopy, transmission electron microscopy, energy-dispersive X-ray spectroscopy, X-ray diffraction, dynamic light scattering, and zeta potential. The physicochemical characterization exhibits the formation of spherical, well-arranged, crystalline structures, with sizes of 12–45 nm. The nanocomposites demonstrated exceptional hemocompatibility (1.7% hemolysis at 1,000 μg·mL−1). Antioxidant evaluations showed potent activity in both the extract (2,2-diphenyl-1-picrylhydrazyl [DPPH] IC50: 25.66 μg·mL−1; 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) [ABTS] IC50: 33.36 μg·mL−1) and CZ nanocomposite (DPPH IC50: 42.71 μg·mL−1; ABTS IC50: 47.34 μg·mL−1), with the nanocomposite exhibiting superior total antioxidant capacity (394.08 AAE μg·mg−1) and ferric reducing antioxidant power (362.57 AAE μg·mg−1) values. The CZ nanocomposite demonstrated stronger anti-inflammatory activity (COX-1 IC50: 22.72 μg·mL−1; COX-2 IC50: 33.03 μg·mL−1) than the extract (COX-1 IC50: 205.54 μg·mL−1; COX-2 IC50: 397.18 μg·mL−1). Antimicrobial results revealed that the fungal extract exhibited superior inhibitory zones against Staphylococcus aureus (30 mm), Bacillus subtilis (29 mm), Escherichia coli (30 mm), and Salmonella typhi (25 mm), exceeding gentamicin performance, while the CZ nanocomposite showed exceptional activity against Enterococcus faecalis (32 mm). For Candida species, the nanocomposite demonstrated superior inhibition against Candida albicans (35 mm, minimum inhibitory concentration [MIC] 7.8 μg·mL−1), while the extract showed better activity against Candida tropicalis (27 mm, MIC 15.62 μg·mL−1). This work highlights the potential of Red Sea fungi as sources of bioactive compounds and green synthesis of functional nanomaterials for pharmaceutical applications.
1 Introduction
Marine environments contain vast numbers of unexplored fungal species that survive extreme oceanic conditions by developing distinct biochemical pathways [1]. The Red Sea’s high salinity and temperature variations create favorable conditions for isolating novel strains that produce enhanced bioactive compounds [2]. These oceanic microbes, especially those inhabiting distinctive underwater realms, yield metabolites valuable for drug development [3]. Marine fungi produce structurally complex secondary metabolites with strong antimicrobial, anticancer, and antioxidant activities through self-produced protective compounds that help them withstand extreme environments [4].
Research from the World Data Centre for Microorganisms (WDCM) reported that among the 378 known Aspergillus species, 180 possess distinct pharmaceutical and industrial importance [5]. Among the prominent species, Aspergillus fumigatus, Aspergillus terreus, Aspergillus niger, and Aspergillus flavus excel at biosynthesizing bioactive natural products [6]. The metabolic repertoire spans multiple chemical families: terpenes, alkaloids, coumarins, quinones, and flavonoids [7]. A. niger grows reliably under different conditions and produces various bioactive compounds, making it suitable for biotechnological work [8]. Analysis of Aspergillus extracts has uncovered several metabolites, with A. niger yielding both common organic acids (fumaric acid, succinic acid, 4-hydroxybenzoic acid) and previously undocumented compounds like 4-(2-hydroxyethyl) phenol and N-[2-(4-hydroxyphenyl) ethyl] acetamide [8,9]. The extracts’ rich composition of phenolics, hydrocarbons, and phthalate derivatives underlies their therapeutic potential, exhibiting inflammation-suppressing, tumour-fighting, microbe-inhibiting, and radical-scavenging capabilities [10].
Modern nanotechnology focuses on environmentally safe synthesis methods that avoid toxic chemical reducing agents while maintaining particle stability and biological activity [11]. Fungal metabolite routes for NPs synthesis offer rapid particle formation through natural enzymatic processes, cost-effective production without specialized equipment requirements, and inherent biocompatibility due to organic compound integration [12]. The production pathway harnesses fungal metabolism to generate these particles without toxic chemicals, yielding stable compounds enriched with natural molecules that readily integrate into biological systems [13].
The convergence of marine fungal biotechnology and green nanotechnology presents unprecedented opportunities for developing next-generation antimicrobial therapeutics [14]. Current limitations in conventional antibiotic efficacy, coupled with the urgent need for biocompatible treatment modalities, necessitate innovative approaches that harness natural biological processes for therapeutic applications [15,16]. Fungal-derived NPs represent a breakthrough in pathogen control studies, combining green synthesis methods with robust pathogen-fighting efficacy that surpasses conventional antibiotic treatments [17]. Their multi-targeted approach to inhibit infectious microbes from breaking down cell structures to disrupting vital processes becomes even more formidable when paired with standard antimicrobial medications, offering renewed hope against treatment-resistant infections [18].
The green synthesis of CuO/ZnO-NPs has positioned them at the forefront of medical innovation, marked by superior cellular compatibility and minimal toxic effects [19]. Testing against resistant bacterial strains revealed their potent antimicrobial action, particularly in inhibiting Neisseria gonorrhoeae and Staphylococcus aureus proliferation [20] while sparing human cells from damage, as confirmed through elevated IC50 measurements [21]. Beyond their antimicrobial potential, these CuO/ZnO-NPs exhibit multifunctional potential in healthcare applications. Their incorporation into drug delivery systems enhances therapeutic efficacy while minimizing adverse effects [22]. Their natural radical-scavenging traits further aid in treating oxidative stress-related conditions. Studies on nanofluid movement in blood vessels showed improved transport mechanisms, especially in complex vessel shapes, where electroosmotic forces help distribute particles more effectively [23,24]. The combination of these biological features, paired with proven safety data, positions biogenic CuO/ZnO-NPs as crucial elements in therapeutic interventions [25].
This study endeavored to investigate the bioactive potential of marine fungi isolated from the Red Sea, with particular emphasis on A. niger AMG31. The research encompassed the isolation and molecular characterization of the fungal strain, followed by comprehensive chemical profiling of its metabolites using fourier transform infrared spectroscopy (FT-IR), gas chromatography–mass spectrometry (GC–MS), and high-performance liquid chromatography (HPLC) analyses, alongside quantitative assessment of key phytochemical constituents, including phenolics, flavonoids, tannins, saponins, and alkaloids. The work further explored the biosynthesis of copper zinc oxide nanocomposite (CZ nanocomposite) using the fungal extract as a green reducing agent. The study evaluated the biological activities of both the crude fungal extract and the mycosynthesized CZ nanocomposites, examining their antioxidant properties through multiple assays (2,2-diphenyl-1-picrylhydrazyl [DPPH], 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) [ABTS], total antioxidant capacity [TAC], and ferric reducing antioxidant power [FRAP]), anti-inflammatory effects via COX-1 and COX-2 inhibition, and antimicrobial efficacy against various pathogenic bacteria and Candida species. Additionally, the research assessed the hemocompatibility of the synthesized CZ nanocomposites to determine their potential therapeutic applications.
2 Materials and methods
2.1 Isolation and characterization of marine fungi from the Red Sea
Seawater samples were collected from the Red Sea and diluted with 0.9% saline solution as a preparatory step. The diluted samples were then plated on potato dextrose agar and prepared with seawater using the spread plate technique to disperse the sample evenly. Following inoculation, the plates were incubated and periodically checked for the growth of fungal colonies starting at 14 days post-plating. Additionally, observation of morphological characteristics and microscopy facilitated identifying and enumerating the isolated marine fungal groups.
The 18S rRNA sequencing was performed to determine the genetic classification of the fungal isolates. BLAST analysis compared these sequences with existing entries in the GenBank database to identify matching patterns. Further examination of the aligned sequences between the isolates and database entries revealed the underlying phylogenetic associations [26].
To investigate bioactive compounds produced by the isolated fungi, liquid cultures were initiated by inoculating yeast extract broth and incubating for 14 days at 25°C. After this growth period, filtration through the Whatman filter paper separated the fungal biomass from the surrounding culture supernatant. The recovered culture supernatant was subjected to liquid-liquid extraction with ethyl acetate to partition secreted fungal molecules into the organic phase. Finally, the ethyl acetate fraction was concentrated via rotary evaporation to obtain a concentrated extract enriched in extracellular compounds synthesized by the marine fungi during cultivation [27].
2.2 Chemical characterization of the fungal extract
2.2.1 Functional group profiling of the fungal extract via FT-IR spectroscopy
FTIR spectroscopy was employed to analyze the functional groups in the fungal extract. The extract was dried at 45°C for 24 h in a desiccator and then thoroughly blended with KBr. A thin disk made from this extract-KBr mixture was analyzed using a Nicolet 6700 FTIR spectrometer (Thermo Fisher Scientific) within the 400–4,000 cm−1 range [28], allowing the detection of characteristic absorption bands corresponding to the extract’s functional groups.
2.2.2 Chemical profiling of the fungal extract using GC–MS technique
The chemical composition was analyzed using a Thermo Scientific TRACE GC1310-ISQ mass spectrometer equipped with a TG-5MS capillary column (30 m × 0.25 mm × 0.25 μm film). The oven temperature program was 35°C initially, increased from 3°C·min−1 to 200°C (held at 3 min), then to 280°C at 3°C·min−1 (held for 10 min). Injector and MS transfer line temperatures were 250 and 260°C, respectively. Helium was the carrier gas at a 1 mL·min−1 flow rate. A 3 min solvent delay was set, and 1 μL of diluted samples was injected in split mode via the AS1300 autosampler. EI mass spectra were acquired at 70 eV over m/z 40–1,000 in full scan mode, with an ion source at 200°C [29]. Components were identified by matching retention times and mass spectra to Wiley 09 and NIST 11 databases.
2.2.3 HPLC analysis
HPLC analysis was performed using an Agilent 1260 system with a Zorbax Eclipse Plus C8 column (4.6 mm × 250 mm, 5 μm particle size). The mobile phase comprised H2O (solvent A) and 0.05% CF3COOH in CH3CN (solvent B). A linear gradient elution protocol was implemented as follows: 0–1 min, 82% A; 1–11 min, 82–75% A; 11–18 min, 75–60% A; 18–22 min, 60–82% A; 22–24 min, 82% A. The flow rate was 0.9 mL·min−1 throughout the 24 min run time [30]. Detection was done using a variable wavelength detector set at 280 nm. The injection volume for samples was 5 μL, and the column temperature was maintained at 40°C.
2.2.4 Spectrophotometric determination of flavonoid content in fungal extracts
One milliliter of the fungal extract was dissolved in 2 mL of CH3OH in a 10 mL volumetric flask. 5% NaNO3, 5% NaOH, and 7% AlCl3 solutions were prepared using H2O in 25 mL volumetric flasks. Two hundred microliters of the extract was taken in a sealed glass vial, and 75 μL of 5% NaNO3 was added and left for 5 min at room temperature. Then, 1.25 mL of AlCl3 and 0.5 mL of NaOH were added to each vial. The mixture was then sonicated and incubated for 5 min at room temperature. After incubation, the absorbance of all working and standard solutions was measured against a methanol blank at 510 nm [31]. The flavonoid content of the extracts was estimated using a quercetin standard calibration curve, and the results were expressed as micrograms of quercetin equivalent (Qu) per 1 g of dry extract.
2.2.5 Total phenolic quantification by the Folin–Ciocalteu method
The total phenolic content of the extracts was determined using a colorimetric assay with the Folin–Ciocalteu reagent. One milliliter of the fungal extract was dissolved in 2 mL of CH3OH. 500 μL aliquots of the extract solution were mixed with 2.5 mL of 10-fold diluted Folin–Ciocalteu reagent and 2.5 mL of 75 g·L−1 Na2CO3 solution. The mixtures were vortexed for 10 s and then incubated at 25°C for 2 h. After incubation, the absorbance was measured at 765 nm against a reagent blank [32]. The total phenolic content was expressed as gallic acid equivalent milligrams per gram.
2.2.6 Quantification of fungal tannin content
The tannin content was quantified using tannic acid as the reference compound. Four hundred microliters of the extract was mixed with 3 mL of 4% vanillin solution in CH3OH and 1.5 mL of concentrated HCL. After incubating this reaction mixture for 15 min, the absorbance was measured at 500 nm [33].
2.2.7 Vanillin-H2SO4 colorimetric determination of fungal saponins
One gram of powdered fungal extract was mixed with 30 mL CH3OH in a 100 mL flask, left for 30 min pre-leaching at ambient temperature, then rapidly cooled in an ice bath and filtered through a 0.45 μm membrane to obtain a clear extract. The extract was transferred to tubes and evaporated to dryness at 65°C, then 0.5 mL 4% vanillin in C2H5OH and 2.5 mL 72% H2SO4 were added, tubes covered, vortexed, incubated at 60°C for 15 min, and then cooled 5 min [34]. Absorbance was measured at 560 nm using a Biosystem 310 spectrophotometer after zeroing with a blank. The standard curve is constructed by plotting absorbance vs concentration.
2.2.8 Spectrophotometric determination of total alkaloids using BCG reagent
Three grams of the fungal extract was dissolved in 2 N HCl and filtered. One milliliter of this solution was transferred to a separatory funnel and washed three times with 10 mL CHCL3. The pH was adjusted to neutral with 0.1 N NaOH. Then, 5 mL of BCG (bromocresol green) solution and 5 mL phosphate buffer were added. The mixture was shaken, and the complex was extracted with 1, 2, 3, and 4 mL CHCL3 by vigorous shaking. The chloroform extracts were collected in a 10 mL volumetric flask and diluted to volume with chloroform. The absorbance of the complex in chloroform was measured at 470 nm [35].
2.2.9 Biogenic CZ nanocomposite synthesis from A. niger AMG31
Fungal metabolites were obtained by cultivating a single colony of A. niger AMG31 in Czapek Dox medium under controlled conditions (27°C, week-long incubation with constant agitation at 150 rpm). The resulting growth underwent filtration through Whatman No. 1 paper and multiple sterile water rinses. The cleaned biomass (7 g) was then suspended in 100 dH2O to achieve a final concentration of 70 g·L−1 (0.07 g·mL−1) and further incubated for a day at 27°C. The bimetallic CZ nanocomposite synthesis was initiated by combining 20 mL of the fungal extract with 200 mL of ZnSO4·7H2O solution (2 g, 0.0348 M stock concentration, 0.0218 M final concentration), followed by pH adjustment to 10 using NaOH. Maintaining rapid agitation, the blend was warmed until a white coloration emerged (30 min). Subsequently, CuSO4·5H2O (2 g in 100 mL, 0.0801 M stock concentration, 0.0250 M final concentration)was carefully dispensed into the heated mixture (85°C) over 5 min, transforming it to dark green. The mixture was allowed to cool to room temperature (25°C) over 15 min before post-2 h agitation, and then, the suspension underwent sequential purification via dH2O rinsing (three cycles) and high-speed centrifugation (10,000 rpm, 15 min). Extended thermal treatment (50°C, 48 h) yielded the final CZ nanocomposite [36].
2.3 Chemical characterization of the mycosynthesized CZ nanocomposite
Chemical bonding and functional group analysis of the biosynthesized CZ nanocomposite were performed using a Cary-660 FT-IR spectrophotometer. The analysis involved preparing KBr pellets by compressing a mixture of 10 mg CZ nanocomposites with potassium bromide powder. The resulting pellets were scanned across a 400–4,000 cm⁻¹ wavenumber range to obtain their vibrational spectra [37]. X-ray diffraction (XRD) patterns of fungal-derived CZ nanocomposites were recorded on a PANalytical-X’Pert-Pro-MRD diffractometer using CuKα radiation (λ = 1.54 Å). Diffraction data collected between 10° and 80° (2θ) at 40 kV and 30 mA revealed the crystalline nature of the synthesized particles [38].
Transmission electron microscopy (TEM) analysis (JEOL Ltd-1010, Tokyo, Japan) determined the size distribution and morphology of the CZ nanocomposite. The powder underwent ultrasonication in water, followed by drop-casting onto carbon-coated TEM grids and air-drying before microscopic examination [39]. Size distribution measurements of CZ nanocomposite in colloidal suspension were obtained through dynamic light scattering analysis using Malvern Nano-ZS (Malvern Ltd., UK) [40]. The nanoparticles dispersed in MilliQ water eliminated signal interference during measurements. Surface charge characteristics were determined using Malvern Nano-ZS Zeta-sizer under identical conditions [41].
2.4 Hemocompatibility evaluation of CZ nanocomposite
Red blood cell (RBC) suspension was prepared and washed with 150 mM NaCl. The samples were centrifuged at 2,500 rpm for 10 min and resuspended in phosphate-buffered saline (pH 7.4). CZ nanocomposite concentrations between 50 and 1,000 μg·mL−1 underwent testing. The RBC solution was mixed with CZ nanocomposite serial dilutions to reach a 1 mL final volume. After 60 min of incubation at 37°C and centrifugation at 2,500 rpm for 15 min, the supernatant was collected for absorbance measurements at 546 nm [42]. Isotonic solutions and full hemolysis served as negative and positive controls.
2.5 Antioxidant assessment of the fungal extract and CZ nanocomposite
2.5.1 DPPH evaluation
The DPPH radical scavenging assay was performed as follows. A 0.1 mM DPPH solution was prepared in ethanol. One milliliter of this DPPH solution was added to 3 ml of the fungal extract solution and CZ nanocomposite at varying concentrations ranging from 3.9 to 1,000 μg·mL−1. Each mixture was shaken and incubated at 37℃ for 30 min. After incubation, the absorbance was measured at 517 nm using a UV-VIS spectrophotometer (Milton Roy) [43].
2.5.2 ABTS˙+ scavenging activity
The ABTS radical cation (ABTS˙+) was generated by reacting 7 mM ABTS stock solution with 2.45 mM K2S2O8 and allowing the mixture to stand at 25°C in the dark. The ABTS˙+ radical scavenging activity of the fungal extract and the CZ nanocomposite was determined by adding 0.07 mL of the fungal extract and CZ nanocomposite separately to 3 mL of diluted ABTS˙+ solution. After incubating for 6 min, the absorbance was measured at 734 nm using a spectrophotometer [44]. The radical scavenging capacity was expressed relative to the standard antioxidant gallic acid as the reference:
2.5.3 TAC assessment
The total antioxidant capacity (TAC) of the fungal extract and the CZ nanocomposite was determined using the phosphomolybdenum method, where 0.5 mg·mL−1 of the extract was mixed with a reagent containing 0.6 M H2SO4, 28 mM NaH2PO4, and 4 mM ammonium molybdate. A blank solution with only the reagent was prepared. The mixtures were incubated at 95°C for 150 min and cooled to 25°C, and absorbance was measured at 630 nm using a microplate reader [44]. Results were expressed as ascorbic acid equivalent (AAE) in µg·mg−1.
2.5.4 FRAP evaluation assay
The potassium ferricyanide/trichloroacetic acid method was employed and adapted for a microplate assay to evaluate the influence of solvent polarity on the fungal extract and CZ nanocomposite reducing power. In Eppendorf tubes: 40 μL of the fungal extract solution was combined with 50 μL 0.2 M Na2HPO4·2H2O buffer, 50 μL 1% K3Fe(CN)6, and 50 μL 10% C2HCl3O2. The mixtures were centrifuged at 3,000 rpm for 10 min. 160 μL supernatant from each tube was transferred to a 96-well plate and mixed with 33.3 μL 1% FeCl3. Absorbance was measured at 630 nm using a microplate reader [45]. DMSO was the negative control, and 1 mg·mL−1 ascorbic acid was the positive control. Results expressed as AAE in µg·mg−1.
2.6 Anti-inflammatory evaluation of the fungal crude extract and CZ nanocomposite
The in vitro ability of the fungal extract and the CZ nanocomposite to inhibit COX-1 and COX-2 isoenzymes was determined using COX-1 (catalogue number k548) and COX-2 (catalogue number k547) inhibitor screening assay kits (Biovision, USA), respectively, following the manufacturer’s instructions [46]. The extract was dissolved in DMSO and tested at concentrations ranging from 1,000 to 0.5 μg·mL−1 in a final volume of 1 mL [47]. Celecoxib was used as a positive control for both COX-1 and COX-2 inhibition assays. The samples were tested in triplicate at 12 different concentrations.
2.7 Antibacterial activity of the fungal extract and CZ nanocomposite
The antibacterial potential of the fungal extract was investigated against various bacterial strains using the agar well diffusion method on Mueller–Hinton agar. The tested bacteria included Gram-positive Bacillus subtilis (ATCC 6633), S. aureus (ATCC 6538), Enterococcus faecalis (ATCC 29212), and Gram-negative Escherichia coli (ATCC 8739), Klebsiella pneumoniae (ATCC 13883), and Salmonella typhi (ATCC 6539). The Candida species tested were Candida albicans (ATCC 10221) and Candida tropicalis (ATCC 66029), and their results were compared to those of fluconazole as the antifungal control drug. The fungal extract and gentamicin (control) were dissolved in DMSO at a 10 mg·mL−1 concentration. Wells were loaded with 10 mg·mL−1 units of fungal extract [44]. After incubation, the diameters of inhibition zones around the wells were measured. Minimum inhibitory concentration (MIC) and minimum bactericidal concentration (MBC) were further evaluated following Clinical and Laboratory Standards Institute (CLSI) guidelines [48].
3 Results and discussion
3.1 Isolation, screening, and molecular classification of Red Sea fungal strains
Fungal sampling from Red Sea sediments yielded eight isolates through dilution methods. Cross-streak tests identified AMG31 as the strain exhibiting superior characteristics to other recovered isolates. Isolate AMG31’s microscopy revealed the hallmark features of A. niger, with a prominent dark brown to black spherical vesicle. The vesicle exhibits dense radiating chains of conidia around its surface, forming a distinctive globose head. The conidiophore appears smooth-walled and unbranched, extending from the base of the vesicle. The conidia appear as small, spherical structures, displaying the characteristic dark pigmentation typical of A. niger (Figure 1a).

(a) Microscopic examination of A. niger AMG31. (b) Phylogenetic association tree of AMG31 with closely related Aspergillus species.
Molecular identification through sequence analysis (accession PP493930) confirmed the morphological identification, positioning isolate AMG31 within a distinct clade of the A. niger complex in the phylogenetic tree. The phylogenetic analysis reveals close evolutionary relationships with A. niger isolate Gharib 11 (PP359563) and isolate NAS-A102 (PQ606657), as indicated by their phylogenetic clustering and bootstrap support values (Figure 1b).
3.2 Chemical characterization of the fungal crude extract
3.2.1 FT-IR
The FT-IR analysis of the fungal extract revealed the presence of several key functional groups, as indicated by firm peaks in the spectrum. A strong, broad peak at 3,300 cm−1 suggests the existence of hydroxyl groups from alcohols or carboxylic acids, while a strong peak at 1,709 cm−1 is attributed to carbonyl groups from carboxylic acids, esters, or ketones. A strong peak at 1,033 cm−1 also supports C–O stretching from alcohols or esters (Figure 2, Table 1). These findings indicate that the fungal extract contains compounds with various functional groups, including alcohols, carboxylic acids, esters, and ketones.

FT-IR spectrum of A. niger AMG31 crude extract.
FT-IR analysis of the fungal extract: Observed peak wavelengths, functional groups, and bond types
Wavelength (cm−1) | Functional group | Bond type | Relative intensity |
---|---|---|---|
3,300 | O–H (alcohol, carboxylic acid) | Stretching | Strong, broad |
2,918 | C–H (alkyl groups) | Stretching (CH2) | Medium |
2,850 | C–H (alkyl groups) | Stretching (CH2) | Medium |
1,709 | C═O (carboxylic acid, ester, ketone) | Stretching | Strong |
1,450–1,200 | C–H (alkyl groups), C–O (alcohol, ester) | Bending (C–H), Stretching (C–O) | Medium |
1,033 | C–O (alcohol, ester) | Stretching | Strong |
719 | −(CH2)n-(long alkyl chain | In-plane bending | Weak |
3.2.2 GC–MS
Looking at the GC–MS analysis results of the fungus’s secondary metabolites, several significant compounds stand out. The most abundant compound is p-cresol, 2,2′-methylenebis[6-tert-butyl], comprising 24.17% of the total area. Following this, (Z)-13-docosenamide represents another major component at 16.18%. The analysis reveals a substantial presence of phenolic compounds, with phenol, 2,4-bis(1,1-dimethylethyl)-,1,1′,1″-phosphate accounting for 10.75%. The fungus also produces notable amounts of bioactive compounds like pyrrolo[1,2-a]pyrazine-1,4-dione derivatives (5.00%) and γ-sitosterol (4.26%). The presence of various long-chain hydrocarbons, fatty acid derivatives, and steroid-like structures suggests that A. niger AMG31 has diverse biosynthetic capabilities. The detection of glucosylated compounds like 6,8-DI-C-α-glucosylluteolin (3.05%) and various amines, including dimethyl myristamine (3.12%), indicates complex secondary metabolism pathways (Table 2 and Figure 3).
Chemical compounds identified by GC–MS analysis
RT | Compound name | Chemical formula | MW | Area (%) |
---|---|---|---|---|
31.99 | 2-Allyl-5-t-butylhydroquinone | C13H18O2 | 206 | 0.25 |
32.2 | 1-Dodecanamine,N,N-dimethyl | C14H31N | 213 | 1.73 |
36.05 | Hexadecane | C16H34 | 226 | 0.42 |
37.96 | Docosane | C22H46 | 310 | 0.22 |
38.83 | n-Lauryl acrylate | C15H28O2 | 240 | 0.58 |
39.69 | Dimethyl myristamine | C16H35N | 241 | 3.12 |
40.02 | 2,6,10-Trimethyltetradecane | C17H36 | 240 | 0.43 |
40.65 | Cyclo(l-prolyl-l-valine) | C10H16N2O2 | 196 | 0.65 |
43.15 | Octadecane | C18H38 | 254 | 0.46 |
43.58 | Phytan | C20H42 | 282 | 1.01 |
44.67 | Pyrrolo[1,2-a]pyrazine-1,4-dione, hexahydro-3-(2-methylpropyl)- | C11H18N2O2 | 210 | 5.00 |
45.39 | 7,9-Di-tert-butyl-1-oxaspiro(4,5)deca-6,9-diene-2,8-dione | C17H24O3 | 276 | 1.33 |
46.45 | 2,2,3,3,4,4 Hexadeutero octadecanal | C18H30D6O | 274 | 0.32 |
46.77 | Methyl 14-methylpentadecanoate | C17H34O2 | 270 | 0.42 |
48.08 | Hexadecanoate | C16H32O2 | 256 | 1.03 |
40.31 | 1-Eicosanol | C20H42O | 298 | 1.17 |
49.6 | 14-á-H-Pregna | C21H36 | 288 | 0.48 |
50.27 | 2-Acetyl-3-(2-cinnamido)ethyl-7-methoxyindole | C22H22N2O3 | 362 | 0.19 |
51.7 | E,E,Z-1,3,12-Nonadecatriene-5,14-diol | C19H34O2 | 294 | 0.62 |
52.01 | Octadecanal, 2-bromo | C18H35BrO | 346 | 0.26 |
52.33 | N-Benzyl-N-methyl-1-tetradecanamine | C22H39N | 317 | 0.90 |
52.94 | MethyL-9,9,10,10-D4-octadecanoate | C19H34D4O2 | 302 | 0.23 |
55.61 | Docosan-1-OL | C22H46O | 326 | 1.73 |
55.73 | Eicosyl acetate | C22H44O2 | 340 | 1.08 |
56.24 | 17-Pentatriacontene | C35H70 | 490 | 0.33 |
56.65 | N,N-Dimethylpalmitamide | C18H37NO | 283 | 1.00 |
58.11 | 17-Methoxyandrost-4-en-3-one o-methyloxime | C21H33NO2 | 331 | 0.61 |
58.25 | Glycidol oleate | C21H38O3 | 338 | 0.85 |
59.34 | Ethyl 2-[(4-methylphenyl)amino]propanoate | C12H17NO2 | 207 | 1.30 |
61.33 | p-Cresol, 2,2′-methylenebis(6-tert-butyl) | C23H32O2 | 340 | 24.17 |
62.37 | 1-Hexacosene | C26H52 | 364 | 1.92 |
64.28 | 1,25-Dihydroxyvitamin D3, TMS derivative | C30H52O3Si | 488 | 1.31 |
64.68 | 2,3-Dihydroxypropyl palmitate | C32H64O3 | 330 | 1.31 |
64.78 | 6,8-DI-c-á-glucosylluteolin | C27H30O16 | 610 | 3.05 |
65.59 | 9-(2′,2′-Dimethylpropanoilhydrazono)-3,6-dichloro-2,7-bis-[2-(diethylamino)-ethoxy]fluorene | C30H42Cl2N4O3 | 576 | 1.82 |
66.51 | E-10,13,13-Trimethyl-11-tetradecen-1-ol acetate | C19H36O2 | 296 | 1.18 |
69.78 | 7,8-Epoxylanostan-11-ol, 3-acetoxy | C32H54O4 | 502 | 0.92 |
71.58 | 13-Docosenamide, (Z)- | C22H43NO | 337 | 16.18 |
80.13 | (R)-2,7,8-Trimethyl-2-((3E,7E)-4,8,12-trimethyltrideca-3,7,11-trien-1-yl)chroman-6-ol | C28H42O2 | 410 | 1.89 |
80.39 | Arabinitol, pentaacetate | C15H22O10 | 362 | 0.46 |
83.08 | ç-Sitosterol | C29H50O | 414 | 4.26 |
83.38 | (E)-24-Propylidenecholesterol | C30H50O | 426 | 2.66 |
83.97 | 3-Hydroxyspirost-8-en-11-one | C27H40O4 | 428 | 0.39 |
91.58 | Phenol, 2,4-bis(1,1-dimethylethyl)-,1,1′,1″-phosphate | C42H63O4P | 662 | 10.75 |

GC–MS analysis of chemical compounds of A. niger’s secondary metabolites.
3.2.3 HPLC
The HPLC analysis of the fungal crude extract revealed Hesperetin as the predominant secondary metabolite, reaching 80,471.56 μg·g−1, followed by Rosmarinic acid at 8,396.08 μg·g−1. Daidzein emerged as the third most concentrated compound at 3,172.03 μg·g−1, while naringenin and gallic acid showed substantial levels at 2,630.41 and 2,415.22 μg·g−1, respectively. Chlorogenic acid maintained a notable presence at 1,751.18 μg·g−1, whereas compounds like methyl gallate and ellagic acid appeared in trace amounts at 33.25 and 77.80 μg·g−1 (Figure 4 and Table 3).

Secondary metabolite profiling of A. niger AMG31 by HPLC.
Quantification of bioactive metabolites in A. niger AMG31
A. niger AMG31 crude extract | |||
---|---|---|---|
Area | Conc. (µg·mL−1) | Conc. (µg·g−1) | |
Gallic acid | 546.10 | 48.30 | 2415.22 |
Chlorogenic acid | 269.92 | 35.02 | 1751.18 |
Catechin | 0.00 | 0.00 | 0.00 |
Methyl gallate | 13.20 | 0.66 | 33.25 |
Coffeic acid | 279.16 | 21.60 | 1080.12 |
Syringic acid | 94.40 | 6.90 | 345.19 |
Pyro catechol | 0.00 | 0.00 | 0.00 |
Rutin | 59.91 | 8.84 | 441.88 |
Ellagic acid | 15.58 | 1.56 | 77.80 |
Coumaric acid | 85.47 | 3.04 | 152.08 |
Vanillin | 235.66 | 8.76 | 437.90 |
Ferulic acid | 62.60 | 3.64 | 181.82 |
Naringenin | 575.56 | 52.61 | 2630.41 |
Rosmarinic acid | 1566.17 | 167.92 | 8396.08 |
Daidzein | 1131.20 | 63.44 | 3172.03 |
Querectin | 33.08 | 4.47 | 223.27 |
Cinnamic acid | 715.91 | 12.82 | 641.03 |
Kaempferol | 227.42 | 14.35 | 717.28 |
Hesperetin | 32735.20 | 1609.43 | 80471.56 |
3.2.4 Total phytochemical screening
The phytochemical screening of metabolites extracted from A. niger revealed a diverse profile of bioactive compounds, with total phenolic content dominating at 55.517 mg(gal)·g−1, representing the most abundant metabolite in the extract. Following this, flavonoids were found to be the second most prevalent compound at 28.757 mg(QuE)·g−1, while tannins showed a moderate presence of 18.650 mg(TanE)·g−1. The extract contained saponins at 12.913 mg(AE)·g−1, whereas alkaloids had the lowest concentration at 5.570 mg·g−1 (Table 4).
Quantitative assessments of the fungal crude extract phytochemicals
A. niger AMG31 crude extract | |
---|---|
Phytochemical screening | Mean ± SD |
Flavonoid conc., mg(QuE)·g−1 | 28.757 ± 0.232 |
Total phenolic conc., mg(gal)·g−1 | 55.517 ± 0.909 |
Total tannins conc., mg(TanE)·g−1 | 18.650 ± 0.450 |
Total saponin content conc., mg AE·g−1 | 12.913 ± 0.301 |
Total alkaloid, mg·g−1 | 5.570 ± 0.305 |
3.3 Characterization of mycosynthesized CZ nanocomposite
3.3.1 FT-IR
The different groups in fungal metabolites and their role in the biofabrication of CZ nanocomposite were detected using FT-IR (Figure 5). As shown, the peaks of the nanocomposite were shifted, decreased in intensity, or increased after coating with fungal metabolites. Moreover, some peaks have disappeared. The strong peak at a wavenumber of 3,320 cm−1 signifies the NH of the primary amide, which overlaps with the OH group [49]. This peak was shifted at 3,400 and 3,570 cm−1 wavenumbers after CZ formation. The medium peaks at 3,055 and 3,010 cm−1 indicate the stretching CH of alkene, whereas the broad peak at 2,670 cm−1 refers to the stretching OH group of carboxylic acid [50]. The peak at 2,315 cm−1 and shifted to 2,365 cm−1 after CZ formation refers to the presence of carbon dioxide (CO2) in the air [51]. The strong peak at 2,185 cm−1 corresponds to the stretching S–C≡N of thiocyanate, whereas weak peaks at 1,800 and 1,870 signify the bending C–H of aromatic compounds [52,53]. The peak at 1,625 cm−1 (shifted to 1,640 cm−1 after CZ nanocomposite formation) related to the stretching C═N for amides and C═O overlapped by N–H of amines [54]. Peaks in the ranges of wavenumbers of 1,250–1,550 cm−1 (five peaks) in the fungal metabolites could be related to the bending C–H of alkane or aldehyde, and stretching N–O of nitro compounds [54].

FT-IR analysis for fungal extract vs CZ nanocomposite showing different functional groups.
These peaks decreased to only two peaks at 1,385 and 1,485 cm−1 after fabrication of the CZ nanocomposite. Moreover, the 1,010–1,210 cm−1 peaks correspond to the stretching C–N of amine. After formation of CZ, the peak at 1,135 cm−1 refers to the stretching vibration of O–H of H2O molecules in the lattice of Cu–Zn–O nanocomposite [55]. The peaks in the ranges of 700–900 cm−1 in fungal extract are related to the bending C═C of alkene [56]. The intensity and number of these peaks are decreased after forming the CZ nanocomposite. The peaks at 695, 620, and 425 cm−1 formed after CZ fabrication are related to the successful formation of Cu–O–Zn [57].
3.3.2 XRD
The nature of the fabricated nanocomposite, whether crystalline or amorphous, was detected using XRD investigation. As shown, the presence of sharp and clearance peaks for Cu and Zn in the XRD analysis indicates the successful doping and formation of a crystallographic Cu/ZnO nanocomposite (Figure 6). Bragg’s peaks at 2 theta degree of 31.8, 34.66, 36.47, 47.71, 56.85, 62.93, 66.64, 68.2, and 69.3 corresponding to (100), (002), (101), (102), (110), (103), (200), (112), and (201) confirmed successful formation of crystallographic hexagonal wurtzite ZnO structure according to JCPDS file 05-0664 [58]. On the other hand, diffraction peaks of (110), (111), (−202), (−113), (310), and (022) at 2 theta degree of 32.9°, 36.47°, 49.2°, 61.22°, 66.64°, and 68.2°, respectively, confirmed formation of crystallographic CuO based on JCPDS card 80-1916 [41]. The results are compatible with different published investigations for green synthesis of crystalline structure of Cu/ZnO nanocomposite [57,59]. The presence of other peaks in XRD analysis could be related to the fungal capping agent, which coated the surface of the nanocomposite and increased its stability.

XRD analysis shows Bragg’s diffraction peaks for CuO and ZnO and confirms the crystalline structure of the nanocomposite.
The average crystallite size for Cu/ZnO nanocomposite was calculated using the Debye–Scherrer equation as follows:
where 1.54 and 0.9 are the wavelength of X-ray and Debye–Scherrer constant, respectively, whereas β and θ are the half maximum intensity and peak Bragg’s angle, respectively.
The overall average CZ crystallite size was calculated to be 23 nm due to the overlapping of the dominant XRD peaks for CuO and ZnO. Similarly, the overall crystallite size of Zn/CuO nanoparticles calculated by the equation of Debye–Scherrer was 26.5 nm [55]. The authors calculated the overall crystallite size due to the overlapping peaks of Cu and Zn in XRD analysis. Also, the average crystallite size of Cu doped with ZnO was in the range of 20.8–28.3 nm as calculated using XRD analysis [57].
3.3.3 Morphological and compositional structure
The shape and composition of the as-formed nanocomposite were investigated using TEM and energy-dispersive X-ray spectroscopy (EDX) analysis. These parameters are vital for analyzing nanocomposites’ biological and environmental applications [60]. TEM images exhibit successful fabrication of spherical shape, well-arranged without agglomeration, and have average sizes of 12–45 nm (Figure 7a and b). Sonkar et al. reported successful formation of Cu doped with varied concentrations of ZnO in the sizes of 19–30 nm and showed that the sizes increased by increasing the amount of ZnO [57]. On the other hand, Cao and coauthors successfully formed the spherical shape of CuO doped with ZnO with sizes in the range of 20–130 nm [21]. The authors returned this wide range of sizes due to the varied ZnO shapes after doping with CuO. Penicillium chrysogenum and Commelina benghalensis facilitated Cu/ZnO nanoparticle formation. The resulting particles mainly assume hexagonal and spherical morphologies, while researchers have also observed rod-like and coral reef-type structures. Nanoparticle dimensions span 9.0–59.7 nm, with ZnO/CuO composite materials measuring 35–50 nm in particular investigations [61]. Another study reported that the sizes of monometallic Cu and Zn are similar and small, whereas they increased with agglomeration percentages after forming bimetallic Cu/Zn nanoparticles due to the interactions of van der Waals [62]. Also, the authors reported that the wide applications of mono and bimetallic NPs are affected by surface characteristics of NPs, such as pore diameter, surface area, and charges, rather than the similarity in morphological properties [62].

Parts (a) and (b) are the TEM analysis of CZ nanocomposites, part (c) is the EDX analysis showing the metal composition of the synthesized sample, and part (d) is the DLS and zeta potential analysis.
The current one-pot synthesis design streamlines the process while maintaining reproducibility, addressing efficiency and cost considerations crucial for potential scale-up applications. While plant extract methods typically require 6–24 h for synthesis completion, bacterial approaches involve lengthy cultivation steps and sterile conditions. Our fungal route completed nanocomposite formation in 2 h at 85°C using a simple sequential addition of metal salts. The fungal extract acted as both a reducing agent and a stabilizer, removing the need for additional chemicals. This approach generated 12–45 nm NPs at lower costs and with fewer processing steps than plant extraction or bacterial fermentation methods.
The green synthesis protocol utilizing 7 g A. niger AMG31 biomass in 20 mL fungal extract combined with 200 mL ZnSO4·7 H2O (2 g) and 100 mL CuSO4·5 H2O (2 g) solutions demonstrated reproducible nanocomposite formation through consistent visual transitions (white to dark green) occurring within 30 min and 2 h timeframes at standardized conditions (pH 10, 85°C, atmospheric pressure, 150 rpm agitation). Reproducibility of nanocomposite characteristics is confirmed by uniform particle size distribution (12–45 nm), stable crystalline structure through XRD analysis, zeta potential measurements via DLS with minimal standard deviations across hemocompatibility testing (1.7% hemolysis at 1,000 μg·mL−1). The thermal treatment (50°C, 48 h) produced dry, crystalline nanocomposites maintaining structural integrity throughout experimental procedures, while comprehensive characterization using six analytical techniques provides quality checkpoints supporting batch consistency. The combination of readily available fungal biomass, common laboratory chemicals, mild processing conditions, and short reaction durations indicated favorable scalability parameters.
The chemical composition of the synthesized nanocomposite was detected using EDX analysis. As shown, the main components of the sample were Cu, Zn, and O ions, which localized at bending energies of (0.9, 8.1, and 8.9 keV), (1.08 and 8.6 keV), and (0.5 keV), respectively (Figure 7c). The data obtained are matched with the EDX of synthesized Cu/ZnO bimetallic nanoparticles. The weight and atomic percentages of the ions within the synthesized sample were presented as 61, 17, and 22% and 59.5, 15.5, and 25% for Cu, Zn, and O ions, respectively (Figure 7c). Similarly, the weight percentages of Cu and Zn in the synthesized nanocomposite were represented by 34.7 and 36.5%, compared to the weight percentages of monometallic NPs (Cu = 74.42% and Zn = 77.42%). Also, the weight percentages of O, Cu, and Zn in green synthesized bimetallic CuO–ZnO nanostructure using the extract of Annona muricata were represented as 53.8, 14.3, and 31.9%, respectively [63]. The high Cu weight and atomic percentages compared to Zn could be related to the high reduction efficacy of Cu (0.34) compared to Zn reduction potential (−0.76) as reported previously [64]. This finding could be because the synthesis of nanocomposites was achieved through mixing two salts simultaneously in the fungal extract under stirring conditions, leading to the reduction of Cu first by the fungal metabolites, followed by Zn.
3.3.4 DLS and zeta potential analysis
The stability of the synthesized nanocomposite in liquid solution was determined by DLS (which detects the size in fluid) and zeta potential (which detects the charge of the nanocomposite surface). As shown, the size obtained by DLS is 60.36 nm (Figure 7d), larger than that of TEM and XRD. The obtained phenomenon may be related to different reasons, such as the sizes obtained by DLS are hydrodynamic (hydrate), whereas TEM sizes are released from the solid or dry sample state. Also, DLS analysis is affected by capping agents or coating substances from the fungal strain, which have a critical role in stability. These substances can increase the DLS sizes. Moreover, the aggregation of samples in liquid media can interfere with DLS analysis, increasing the obtained sizes [65]. Similarly, the DLS size of Cu/Zn bimetallic nanoparticles was 78.8 nm, which is larger than the TEM size (35 nm), and the authors attribute these results to the DLS measuring the hydrodynamic diameter, whereas TEM measures the dry or solid diameter [62].
Moreover, the homogeneity or heterogeneity of the synthesized nanocomposite in the liquid solution can be detected during DLS analysis by measuring the polydispersity index (PDI). The sample distribution based on PDI standard becomes homogenous with narrow sizes at PDI less than 0.4, whereas it becomes heterogeneous at larger PDI (larger than 0.4–1) [66]. In the current investigation, the PDI value was 0.1, which means the high homogeneity distribution of the nanocomposite in the liquid solution with narrow sizes.
The electrochemical equilibrium, surface charge, and stability of nanoparticles in the suspension are mainly understood by zeta potential analysis. The stability degree is detected by the presence of surface charges and the zeta potential value, which becomes unstable at values ranging from ±0 to ±10, moderate at values ranging from ±10 to ±20, and stable at values >±20 [67]. The zeta potential value of the synthesized CZ nanocomposite was −23.5 mV (Figure 7d), which indicates high stability. Also, the presence of one charge (−ve) on its surface keeps the particles far away from each other by electrostatic potential, preventing agglomeration in the solution, and this data is compatible with TEM images. Similarly, the zeta potential analysis of Cu doped with ZnO showed the presence of one charge on its surface with values of −20.1 mV, which indicates the stability of the synthesized nanocomposite [68].
Consistent with our observations, Priestia megaterium-derived nanocomposites showed a zeta potential of −33.4 mV, matching Serratia nematodiphila but exceeding Pseudomonas aeruginosa’s −18.0 mV value [61]. Marine Trichoderma harzianum-based composites exhibited a hydrodynamic diameter of 50.79 nm with a zeta potential of −17.49 mV, demonstrating stable colloidal properties [69].
3.4 Hemocompatibility evaluation of CZ nanocomposite
The hemocompatibility data show the CZ nanocomposite tested at 25–1,000 μg·mL−1 concentration. At 1,000 μg·mL−1, hemolysis reached 1.7%, while 800 μg·mL−1 showed 0.6% hemolysis. The mid-range concentrations of 600, 400, and 200 μg·mL−1 exhibited minimal hemolysis of 0.3, 0.2, and 0.2% respectively. Lower concentrations of 100 and 50 μg·mL−1 displayed 0.2 and 0.1% hemolysis (Figure S1). The lowest tested concentration of 25 μg·mL−1 resulted in no measurable hemolysis (0%), matching the negative control (Table 5). Compared to the positive control’s 100%, the consistently low hemolysis values indicate the nanoparticles’ exceptional compatibility with red blood cells.
Hemocompatibility assessment of the mycosynthesized CZ nanocomposite
Absorbance mean ± SD | Hemolysis (%) | ||
---|---|---|---|
Complete hemolysis (+ve control) | 1.211 ± 0.012 | 100 | |
Isotonic solution (−ve control) | 0.001 ± 0.001 | 0 | |
CZ nanocomposite conc. (μg·mL−1) | 1,000 | 0.075 ± 0.006 | 1.7 |
800 | 0.039 ± 0.001 | 0.6 | |
600 | 0.028 ± 0.002 | 0.3 | |
400 | 0.019 ± 0.002 | 0.2 | |
200 | 0.011 ± 0.003 | 0.2 | |
100 | 0.007 ± 0.002 | 0.2 | |
50 | 0.005 ± 0.002 | 0.1 | |
25 | 0.001 ± 0.001 | 0 |
3.5 Antioxidant activity of the fungal extract and CZ nanocomposite
Antioxidant activity evaluation using a DPPH scavenging assay revealed distinct inhibition patterns between the fungal extract from A. niger and the CZ nanocomposite across multiple concentrations. At the lowest tested concentration of 1.9 μg·mL−1, the fungal extract demonstrated 27% inhibition, while the CZ composite showed 12.6% inhibition. As the concentration increased to 3.9 μg·mL−1, the fungal extract showed 31% scavenging activity, while at 7.8 μg·mL−1, it reached 38.4% inhibition. At 15.6 μg·mL−1, the fungal extract’s inhibition increased to 44.4%, whereas the CZ nanocomposite had 37.5% inhibition. This comparison revealed that while the CZ nanocomposite needed a higher concentration, it achieved a greater inhibition percentage at this point. The fungal extract crossed the 50% inhibition threshold at 31.2 μg·mL−1 with 51.6% scavenging activity, corresponding closely with its calculated IC50 of 25.66 μg·mL−1. In contrast, the CZ nanocomposite reached 54.4% inhibition at 62.5 μg·mL−1 and 63.1% at 125 μg·mL−1, placing its IC50 at 42.71 μg·mL−1. During this concentration range, ascorbic acid had already achieved 74% inhibition at 15.6 μg·mL−1.
At medium concentrations, the fungal extract showed 58.6% inhibition at 62.5 μg·mL−1 and 66.5% at 125 μg·mL−1. The CZ nanocomposite performed slightly better, with 63.1% inhibition at 125 μg·mL−1 and 71.5% at 250 μg·mL−1. By this point, ascorbic acid had reached a significantly higher inhibition of 89.1% at 125 μg·mL−1, with IC50 recorded as 3.26 μg·mL−1. At higher concentrations, the fungal extract achieved 72.2% inhibition at 250 μg·mL−1 and 78.2% at 500 μg·mL−1. The CZ nanocomposite showed comparable results with 71.5 and 79.7% inhibition at 250 and 500 μg·mL−1. At their highest tested concentration of 1,000 μg·mL−1, the fungal extract reached 83.2% inhibition, while the CZ nanocomposite achieved 88.6% (Figure 8).

Dose–response DPPH radical scavenging activity by the fungal extract and CZ nanocomposite compared to ascorbic acid as a reference standard. Results expressed as mean ± SD from triplicate experiments with statistical significance determined by one-way analysis of variance (*p < 0.05, **p < 0.01, ***p < 0.001).
Moving to the ABTS˙ + scavenging assay, at lower concentrations, the fungal extract showed a gradual increase in scavenging activity, starting with 12.1% inhibition at 1.9 μg·mL−1, progressing to 21.6% at 3.9 μg·mL−1, and reaching 29.6% at 7.8 μg·mL−1. The CZ nanocomposite displayed more robust initial activity with 39.9% inhibition at its lowest tested concentration, which progressed to 51.3 and 62.5% at 3.9 μg·mL−1 and 7.8 μg·mL−1. At 15.6 μg·mL−1, the fungal extract demonstrated 40.7% inhibition, while the CZ nanocomposite reached 73.9% inhibition. By 31.2 μg·mL−1, the fungal extract crossed the halfway mark with 51.2% inhibition, corresponding closely with its calculated IC50 of 33.36 μg·mL−1. while CZ nanocomposite IC50 was 47.34 μg·mL−1. Comparatively, gallic acid showed superior activity with 41.8% inhibition at its lowest tested concentration and an IC50 value of merely 2.54 μg·mL−1. The fungal extract continued its progression with 58.4% inhibition at 62.5 μg·mL−1 and 69.9% at 125 μg·mL−1. During this range, the CZ nanocomposite demonstrated 83.2% inhibition and 88.8% at the same tested concentrations. At a maximum concentration tested of 1,000 μg·mL−1, the fungal extract reached 89% inhibition, while the CZ nanocomposite achieved near-complete scavenging with 97.6% inhibition at its highest concentration point, matching exactly the maximum inhibition displayed by gallic acid at 97.6% (Figure 9).

Dose-dependent fungal extract and CZ nanocomposite ABTS˙+ scavenging activity (%) compared to gallic acid standard across the concentration range (1.9–1,000 µg·mL−1). Values shown as mean ± SD (n = 3) with statistical analysis by one-way analysis of variance (*p < 0.05, **p < 0.01, ***p < 0.001).
The TAC assay revealed that the fungal extract possesses a substantial total antioxidant capacity, with a mean value of 334.1 ± 4.3 AAE equivalent µg·mg−1. At the same time, the CZ nanocomposite exhibited significantly higher antioxidant activity at 394.08 ± 4.491 AAE equivalent µg·mg−1, both expressed as AAE. Similarly, the FRAP assay demonstrated the extract’s ability to reduce ferric ions (Fe³⁺) to ferrous ions (Fe²⁺), with a mean value of 252.45 ± 2.57 AAE µg·mg−1. In contrast, the CZ nanocomposite showed markedly superior reducing power at 362.57 ± 4.179 AAE µg·mg−1 (Table 6). These comparative results highlight the enhanced electron-donating capacity of CZ nanocomposite, reinforcing its potential as a more potent antioxidant biomaterial than the fungal extract.
TAC and FRAP results of A. niger AMG31 crude extract
The fungal extract | TAC (equivalent (AAE) µg·mg−1) | FRAP (equivalent (AAE) µg·mg−1) |
---|---|---|
334.1 ± 4.3 | 252.45 ± 2.57 | |
CZ nanocomposite | 394.08 ± 4.491 | 362.57 ± 4.179 |
Data are represented as mean ± SD.
The predominant compound, hesperetin at 80471.56 μg·g−1 concentration, accounts for the robust DPPH (IC50: 25.66 μg·mL−1) and ABTS (IC50: 33.36 μg·mL−1) radical scavenging through electron donation from its hydroxyl groups [70]. While rosmarinic acid (8396.08 μg·g−1) enhances antioxidant capacity by metal chelation and lipid peroxidation prevention through its catechol structure [71]. γ-Sitosterol (4.26% in GC–MS) disrupts bacterial cell walls by altering membrane stability, resulting in antimicrobial effects [72]. The combined phenolic compounds (55.517 mg·g−1) and flavonoids (28.757 mg·g−1) generate the measured TAC (334.1 AAE μg·mg−1) and FRAP (252.45 AAE μg·mg−1) through hydrogen donation and ferric ion reduction mechanisms [73].
Following our data, fungal-derived CuO–ZnO nanocomposites from A. niger exhibit remarkable free radical neutralization capabilities. The biogenic nanocomposite structures efficiently counteract unstable molecules that trigger oxidative stress [74]. A similar composite, but of bacterial origin from Bacillus velezensis spores, enabling rapid oxidation of ABTS into stable ABTS+ radicals without requiring H2O2 [75]. Similarly, a phyto-derived nanocomposite from Allium sativum (garlic) showed significant free radical scavenging capabilities through multiple DPPH, ABTS, and FRAP assays. The authors attributed the antioxidant activity to the heightened electron donation capacity through metal ion synergy, while their structural integrity and dimensions directly influence their radical scavenging performance. These stable, morphologically optimized particles enable sustained antioxidant defense through cooperative metal interactions and enhanced surface reactivity [76].
Comparatively, the fungal metabolite nigerloxin outperformed curcumin in metal ion reduction capacity while showing impressive radical neutralization across multiple testing platforms [77]. Another study reported that A. niger extracts showed enhanced antioxidant potency when cultivated in glucose-enriched media, with internal cellular compounds outperforming secreted metabolites in radical neutralization tests. The fungus produced specific phytochemicals, particularly flavonoids and phenols, that significantly boost its oxidative defense capabilities, underscoring how both growth conditions and extraction methods critically impact its therapeutic potential [78]. The ethyl acetate extract from A. niger showed exceptional radical-neutralizing properties through DPPH evaluation, reaching 95.12% elimination at 1 mg·mL−1 concentration [79]. Another metabolite from the same fungus demonstrated powerful oxidative defense capabilities, inhibiting DPPH radicals beyond 50% while maintaining elevated thiol groups (>29 nmol·μL−1) and superoxide dismutase (SOD) activity (>0.010 units·mL−1) [80].
3.6 Anti-inflammatory activity
Next, the fungal extract and the biosynthesized composite were evaluated for their anti-inflammatory potential by inhibiting COX enzyme activity at various concentrations. The CZ nanocomposite demonstrated substantially superior COX-1 inhibitory activity compared to the fungal extract across all tested concentrations, and this is clearly reflected in the IC50 values: CZ nanocomposite (22.72 ± 0.84 μg·mL−1) showed approximately ninefold greater potency than the fungal extract (205.54 ± 5.98 μg·mL−1), while celecoxib exhibited the most potent inhibitory effect with an IC50 of 6.32 ± 0.74 μg·mL−1 for COX-1 inhibition.
At the lowest tested concentration (0.5 μg·mL−1), both showed minimal COX-1 inhibition, with the fungal extract exhibiting negligible activity (0.67%) compared to the CZ nanocomposite (6.23%). In contrast, celecoxib already demonstrated substantial inhibition (21.38%) at this concentration. When the concentration increased to 2 μg·mL−1, the differences became more pronounced, with the fungal extract showing 3.08% inhibition vs 18.72% for the CZ nanocomposite, while celecoxib reached 36.87%. At 31.2 μg·mL−1, the CZ nanocomposite achieved 58.46% inhibition, surpassing its IC50 value (22.72 μg·mL−1), whereas the fungal extract reached only 20.52% inhibition at this concentration. The fungal extract required a much higher concentration (250 μg·mL−1) to approach its IC50 of 205.54 ± 5.98 μg·mL−1, reaching 56.73% inhibition at this concentration. Notably, at 125 μg·mL−1, the CZ nanocomposite already demonstrated robust inhibition (80.75%), considerably outperforming the fungal extract (37.81%) at the same concentration.
At the highest tested concentration (1,000 μg·mL−1), all three tested groups showed potent inhibitory activity: the fungal extract reached 80.38%, the CZ nanocomposite achieved 94.03%, and celecoxib exhibited the highest inhibition at 96.03%. Even at 500 μg·mL−1, the CZ nanocomposite (91.27%) almost matched celecoxib’s performance (92.74%), whereas the fungal extract trailed considerably with 72.46% inhibition (Figure 10).

Inhibitory effects of fungal extract and CZ nanocomposite on COX-1 enzyme activity across concentration gradient (0.5–1,000 µg·mL−1). Celecoxib served as the positive control standard. Results presented as mean ± SD from triplicate experiments with significance determined by one-way analysis of variance (*p < 0.05, **p < 0.01, ***p < 0.001).
On the other hand, concerning COX-2 inhibition, the enzyme inhibition percentage pattern followed a clear concentration-dependent relationship for all three groups, with substantial differences in their relative potencies. At the lowest tested concentration (0.5 μg·mL−1), the fungal extract exhibited negligible COX-2 inhibition (0.21%), while the CZ nanocomposite showed minimal activity (2.91%). In contrast, celecoxib already demonstrated substantial inhibition (26.37%) at this concentration. Upon increasing to 2 μg·mL−1, the differences became more pronounced, with the fungal extract showing merely 0.75% inhibition vs 13.58% for the CZ nanocomposite, whereas celecoxib reached 39.59%. By raising the concentration of both tested groups to 3.9 μg·mL−1, celecoxib achieved 51.64% inhibition, surpassing its IC50 value (3.73 ± 0.29 μg·mL−1). At the same time, the fungal extract and CZ nanocomposite showed only 1.49 and 21.42% inhibition, respectively. The CZ nanocomposite reached the 50% inhibition threshold at approximately 31.2 μg·mL−1 (49.12%), closely aligning with its IC50 value of 33.03 ± 1.15 μg·mL−1. Meanwhile, at this same concentration, the fungal extract demonstrated minimal inhibition (10.23%), and celecoxib maintained robust activity (72.38%). At 62.5 μg·mL−1, the CZ nanocomposite exhibited considerable inhibition (64.58%), whereas the fungal extract reached only 14.78% inhibition.
At 250 μg·mL−1, the fungal extract achieved just 33.84% inhibition, still considerably below its IC50 value, while the CZ nanocomposite reached 79.43% and celecoxib exhibited 91.63% inhibition (Figure 11). The fungal extract finally approached its IC50 value at 500 μg·mL−1 with 61.29% inhibition, corresponding well with its measured IC50 of 397.18 μg·mL−1. While at the highest tested concentration (1,000 μg·mL−1), all three substances demonstrated strong inhibitory activity: the fungal extract reached 76.82%, the CZ nanocomposite achieved 92.39%, and celecoxib exhibited the highest inhibition at 97.41%.

Concentration-dependent inhibitory effects of fungal extract and CZ nanocomposite on COX-2 enzyme activity across concentration gradient (0.5–1,000 µg·mL−1). Celecoxib served as the positive control standard. Results presented as mean ± SD from triplicate experiments with significance determined by one-way analysis of variance (*p < 0.05, **p < 0.01, ***p < 0.001).
The extract contains rosmarinic acid at 8396.08 μg·g−1, which blocks inflammatory processes by inhibiting NF-κB and directly reducing COX-2 expression [81]. HPLC analysis revealed quercetin (223.27 μg·g−1) and kaempferol (717.28 μg·g−1), both acting through various pathways that include selective COX-2 inhibition and decreased production of inflammatory mediators [82]. GC–MS analysis identified γ-sitosterol at 4.26%, which maintains cellular membrane integrity and regulates phospholipase A2, thereby limiting arachidonic acid release needed for prostaglandin synthesis [72], explaining our observed COX-1 inhibition (IC50: 205.54 μg·mL−1) and COX-2 downregulation (IC50: 397.18 μg·mL−1).
Following our results, Phytobiogenic CuO–ZnO nanocomposites from Dovyalis caffra leaves showed better antioxidant properties than single metal oxides alone. Their scavenging abilities suggest they may inhibit oxidative enzymes, including COX [83]. Plant extract-based copper nanoparticles from Catharanthus roseus and Abutilon indicum displayed anti-inflammatory effects that increased with concentration. This is related to COX inhibition in inflammatory pathways [84]. Also, l-ascorbic acid-synthesized nanocopper demonstrated anticancer and antimicrobial capabilities. These effects are often linked to inhibiting enzymes such as COX, which function in cancer development and inflammation [85].
Copper–zinc nanoparticles show anti-inflammatory effects against COX enzymes through several mechanisms. Beyond COX-2 inhibition, Copper–zinc nanoparticles influence inflammatory cytokines such as TNF-α, IL-1β, and IL-8 [86]. For example, boswellic acid-coated zinc nanoparticles reduce these cytokines, easing inflammation in ulcerative colitis models [86]. Also, copper ions irreversibly inhibit COX-2 activity, lowering the production of inflammatory prostaglandins like PGE2 [87]. These nanoparticles cause less gastrointestinal harm than standard NSAIDs [62,88]. They work by blocking COX-2 and decreasing reactive oxygen species and inflammatory cytokines, as shown in gouty arthritis studies [89].
Furthermore, copper complexes with zinc function like SOD, decreasing superoxide ions and their byproducts that contribute to inflammation [90]. This antioxidant function helps reduce oxidative stress, a key factor in inflammatory processes. These biogenic nanoparticles also improve beneficial immune cell infiltration, including CD8+ T cells, while limiting immunosuppressive cells such as myeloid-derived suppressor cells (MDSCs) [91,92]. Similarly, zinc complexes affect gut microbiota composition, which influences immune responses and decreases the expression of interferon-stimulated response genes, helping resolve inflammation [93].
3.7 Antimicrobial activity
The fungal extract of A. niger and CZ nanocomposite were further evaluated for antimicrobial activity against various G+ve and G−ve bacteria and two different Candida species using the agar well diffusion method. The results were compared to the control antibiotic gentamicin and fluconazole as an antifungal reference drug. MIC and MBC values were determined to evaluate the extract’s potency as well as the nanocomposite’s antimicrobial properties. Overall, the fungal extract exhibited broader spectrum antimicrobial activity than the CZ nanocomposite, demonstrating superior effectiveness against five tested bacterial strains. The CZ nanocomposite, meanwhile, showed its strongest activity specifically against E. faecalis, where it remarkably outperformed both the fungal extract and control gentamicin.
For G+ve bacteria, the fungal extract from A. niger demonstrated remarkable activity against S. aureus with an inhibition zone of 30 mm, surpassing both the CZ nanocomposite (25 mm) and the control gentamicin (25 mm). Similarly, against B. subtilis, the fungal extract produced a substantial inhibition zone of 29 mm, outperforming the CZ nanocomposite (21 mm) yet matching the control gentamicin (29 mm) (Figure 12). For E. faecalis, however, the fungal extract showed moderate effectiveness with an inhibition zone of 22 mm. In contrast, the CZ nanocomposite exhibited superior activity (32 mm), notably exceeding the control gentamicin (27 mm) (Figures S2 and S3).

Comparative antibacterial activity of three treatment agents. (a) Bar graph displaying inhibition zones (mm) (mean ± SD) by fungal extract, CZ nanocomposite, and gentamicin. (b) Representative agar plates showing inhibition zones from treatments where position A is the tested compound (either fungal extract or CZ nanocomposite), position B is blank, and position C is gentamicin.
Turning to G−ve bacteria, the fungal extract exhibited potent activity against E. coli with an inhibition zone of 30 mm, significantly exceeding both the CZ nanocomposite (26 mm) and control gentamicin (24 mm). Against K. pneumoniae, the fungal extract generated an inhibition zone of 22 mm, outperforming the CZ nanocomposite (19 mm) but matching the control gentamicin (22 mm) (Figure 12). Finally, for S. typhi, the fungal extract showed considerable efficacy with an inhibition zone of 25 mm, superior to both the CZ nanocomposite (20 mm) and control gentamicin (21 mm, Figures S2 and S3).
MICs and MBCs were then evaluated for both the fungal extract from A. niger and the CZ nanocomposite against the same bacterial spectrum. The CZ nanocomposite demonstrated superior antibacterial activity, particularly against G+ve bacteria. However, the fungal extract showed better effectiveness against E. faecalis and K. pneumoniae.
For B. subtilis, the CZ nanocomposite exhibited approximately twice the potency of the fungal extract, with an MIC of 15.62 μg·mL−1 compared to the extract’s 31.25 μg·mL−1. Similarly, the MBC values maintain this proportional relationship (31.25 vs 62.50 μg·mL−1). Regarding S. aureus, the nanocomposite again demonstrated enhanced effectiveness with an MIC of 7.8 μg·mL−1, whereas the fungal extract requires 15.62 μg·mL−1. The MBC values followed the same pattern, with the nanocomposite requiring half the concentration (15.62 μg·mL−1) compared to the extract (31.25 μg·mL−1). Interestingly, E. faecalis presents a divergent trend. The fungal extract substantially outperforms the nanocomposite, requiring only 15.62 μg·mL−1 (MIC) compared to the nanocomposite’s 62.5 μg·mL−1. This pattern extends to the MBC values (31.25 vs 125 μg·mL−1) (Table 7).
MIC/MBC antibacterial Profile of CZ nanocomposite vs the fungal extract
CZ nanocomposite | The fungal extract | |||||
---|---|---|---|---|---|---|
Pathogenic bacteria | MIC (µg·mL−1) | MBC (µg·mL−1) | MBC/MIC ratio | MIC (µg·mL−1) | MBC (µg·mL−1) | MBC/MIC ratio |
B. subtilis (ATCC 6633) | 31.25 | 62.50 | 2 | 15.62 | 31.25 | 2 |
S. aureus (ATCC 6538) | 15.62 | 31.25 | 2 | 7.8 | 15.62 | 2 |
E. faecalis (ATCC 29212) | 15.62 | 31.25 | 2 | 62.5 | 125 | 2 |
E. coli (ATCC 8739) | 15.62 | 31.25 | 2 | 7.8 | 15.62 | 2 |
K. pneumoniae (ATCC13883) | 31.25 | 62.50 | 2 | 62.5 | 125 | 2 |
S. typhi (ATCC 6539) | 31.25 | 125 | 4 | 15.62 | 31.25 | 2 |
For the tested G−ve bacterial spectrum, the nanocomposite showed remarkable potency against E. coli with an MIC of 7.8 μg·mL−1, which is notably lower than the fungal extract’s 15.62 μg·mL−1. The corresponding MBC values follow suit (15.62 vs 31.25 μg·mL−1). For K. pneumoniae, the fungal extract performed better, exhibiting an MIC of 31.25 μg·mL−1 compared to the nanocomposite’s 62.5 μg·mL−1. The MBC values reflect this advantage as well (62.50 vs 125 μg·mL−1). With S. typhi, the nanocomposite showed greater effectiveness with an MIC of 15.62 μg·mL−1 compared to the extract’s 31.25 μg·mL−1. The MBC for the nanocomposite (31.25 μg·mL−1) is significantly lower than that of the extract (125 μg·mL−1) (Table 7).
Concerning the anticandidal assessment, the CZ nanocomposite demonstrated superior inhibition activity against C. albicans with a zone of inhibition measuring 35 mm, outperforming both the fungal extract (31 mm) and the standard antifungal fluconazole (30 mm). Conversely, against C. tropicalis, the fungal extract exhibited the strongest inhibitory effect with a 27 mm zone, while the composite produced a 20 mm zone, and fluconazole yielded a 21 mm zone (Figure 13).

Anticandidal activity comparison of three treatment agents. (a) Bar graph displaying inhibition zones (mm) (mean ± SD) by fungal extract, CZ nanocomposite, and fluconazole. (b) Representative agar plates showing inhibition zones from treatments where position A is the tested compound (either fungal extract or CZ nanocomposite), position B is blank, and position C is fluconazole.
Regarding the MIC and MFC values, the CZ nanocomposite demonstrated remarkable potency against C. albicans with an MIC of 7.8 μg·mL−1 and an identical MFC, resulting in an MFC/MIC ratio of 1, whereas the fungal extract showed a higher MIC of 15.62 μg·mL−1 and an MFC of 31.25 μg·mL−1 with an MFC/MIC ratio of 2. Conversely, for C. tropicalis, the fungal extract exhibited greater efficacy with MIC and MFC values both at 15.62 μg·mL−1 and an MFC/MIC ratio of 1, while the CZ nanocomposite required higher concentrations with an MIC of 31.25 μg·mL−1 and MFC of 62.5 μg·mL−1, yielding an MFC/MIC ratio of 2 (Table 8). These results suggest that while the CZ nanocomposite exhibits more vigorous fungicidal activity against C. albicans, the fungal extract proves more effective against C. tropicalis.
MIC/MFC anticandidal profile of CZ nanocomposite vs the fungal extract
Candida species | The fungal extract | CZ nanocomposite | ||||
---|---|---|---|---|---|---|
MIC | MFC | MFC/MIC ratio | MIC | MFC | MFC/MIC ratio | |
Candida albicans (ATCC 10221) | 15.62 | 31.25 | 2 | 7.8 | 7.8 | 1 |
Candida tropicalis (ATCC 66029) | 15.62 | 15.62 | 1 | 31.25 | 62.5 | 2 |
Consistent with our data, biogenic Cu–Zn NPs fabricated by Shewanella oneidensis, its antibacterial inhibitory action was quantified through both diffusion testing and concentration-dependent growth suppression, which was found to be effective against S. aureus and E. coli [94]. Green tea-synthesized copper oxide nanoparticles (GT-CuO NPs) demonstrated remarkable antibacterial efficacy against N. gonorrhoeae, killing 50% of bacteria at just 0.625 μg·mL−1 and 100% at 3.125 μg·mL−1 [95]. Similarly, biogenic Cu–Zn NPs were biologically fabricated, and their antimicrobial assessment revealed superior efficacy with lower MIC values (100 μg·mL−1) and faster killing kinetics (3–3.5 h) against multidrug-resistant S. aureus and E. coli [96]. Similar NPs were synthesized using the methanolic extract of Ocimum sanctum, which demonstrated high antibacterial activity against S. aureus as determined through agar well diffusion, MIC, and MBC assays [97]. Actinomycetes fabricated another nanostructure that suppressed the bacterial growth of Bacillus cereus, showing the highest susceptibility (25.3 mm zone of inhibition) [98].
Biosynthesized Cu–Zn NPs demonstrated potent antifungal action as well against C. albicans with an MIC of 35.5 µg·mL−1, inducing ROS production, while ultrastructural analysis revealed significant nanoparticle accumulation within the cytoplasm, cell wall, and extracellular environment of the fungal cells [99]. Similarly, nanoparticles biosynthesized using Salvia officinalis leaf extract showed potent anticandidal activity, particularly against C. tropicalis (MIC: 40 µg·mL−1, MFC: 80 µg·mL−1), while demonstrating significant synergistic effects when combined with conventional antifungals, showing the highest synergy with terbinafine against C. glabrata and fluconazole against C. albicans [100]. Spherical Cu–Zn nanocomposites were synthesized via biogenic and chemical routes with sizes ranging from 6 to 50 nm, demonstrating 100% growth inhibition against both P. aeruginosa and C. albicans, showing excellent antimicrobial efficacy against these clinically significant pathogens [101].
Biogenic Cu–Zn NPs suppress microbes’ growth by breaking cell walls and disrupting vital functions through several distinct mechanisms. These metallic structures penetrate cell walls, destroying membrane integrity and triggering cell rupture [102]. Also, the particles’ surface properties facilitate molecular alterations to bacterial membranes, compromising their structural stability [103]. Simultaneously, these nanoparticles trigger oxidative damage by producing ROS that attack cellular components such as proteins, lipids, and genetic material, creating an inhospitable environment for microbial survival [104].
Furthermore, the gradual liberation of copper and zinc ions (Cu²⁺ and Zn²⁺) from nanoparticle surfaces interrupts essential metabolic sequences within target organisms [105]. Cu–Zn particles excel at dismantling protective biofilms, removing this crucial bacterial shield and heightening susceptibility to antimicrobial treatments [106]. Not only removing bacterial biofilms, but also these nanoparticles’ interaction with microbial membranes creates abnormal permeability conditions, resulting in cytoplasmic leakage and cellular collapse [107].
Following our current findings, the organic extract from A. niger inhibited suppressed Fusarium oxysporum f. sp. lycopersici proliferation by approximately 51.5% at 80 µg·mL−1 [108]. Likewise, examination of the filtrate crude extract unveiled remarkable fungicidal properties when tested against C. albicans, pointing toward its prospective role in combating fungal pathogens [109]. Additionally, researchers identified a unique antimicrobial polypeptide from A. niger that demonstrated effectiveness at minimal concentrations, requiring merely 8 µg·mL−1 to constrain S. aureus growth, though necessitating a fourfold increase (32 µg·mL−1) to achieve comparable results against drug-resistant MRSA strains [110].
The A. niger AMG31 extract’s notable bactericidal efficacy is rooted in its chemically diverse arsenal targeting multiple bacterial vulnerabilities simultaneously. Phenolic compounds (55.517 mg·g−1), notably hesperetin (80471.56 μg·g−1), disrupt bacterial membranes [111] by intercalating between phospholipids [112], while p-Cresol derivatives (24.17%) denature critical microbial proteins [113]. Meanwhile, flavonoids (28.757 mg·g−1), including daidzein and naringenin, chelate essential metals and interrupt nucleic acid synthesis [114,115], whereas organic acids like rosmarinic acid (8396.08 μg·g−1) acidify bacterial surfaces, compromising membrane integrity [116]. Furthermore, Pyrrolo[1,2-a]pyrazine derivatives (5%) subvert quorum-sensing pathways [117], whereas steroidal compounds such as γ-sitosterol (4.26%) disorganize membrane lipids, particularly effective against Gram-positive species [118].
This multi-pronged chemical approach incorporates amides like (Z)-13-docosenamide (16.18%) functioning as membrane-solubilizing surfactants [119], while tannins (18.650 mg·g−1) irreversibly bind proline-rich proteins [120], and block enzymatic substrate access [121]. Consequently, this sophisticated chemical composition explains the extract’s broad-spectrum efficacy, evidenced by equivalent potency against both E. faecalis and E. coli (MIC 15.62 μg·mL−1), often matching or surpassing gentamicin’s performance. Additionally, the consistent MBC/MIC ratios 2 confirm the extract’s bactericidal rather than bacteriostatic action.
Several future research directions emerge to advance our understanding and applications of marine fungal metabolites and their derived NPs. The initial step involves expanding the screening process to encompass diverse marine fungal strains from various Red Sea depths and locations, which could uncover novel bioactive compounds with unique therapeutic properties. This expansion should be complemented by comprehensive in vivo toxicological assessments of fungal metabolites and BCZ nanoalloy, ensuring their safety profiles for potential pharmaceutical applications. Further research should delve into the underlying mechanistic pathways of these compounds’ antimicrobial and anti-inflammatory effects through molecular docking studies and gene expression analyses, providing deeper insights into their therapeutic mechanisms.
Furthermore, developing optimized formulations for drug delivery systems and investigations into scalable nanoparticle production methods would facilitate the transition from laboratory findings to practical applications. Lastly, exploring the potential synergistic effects between fungal extracts and their synthesized NPs could uncover enhanced therapeutic properties, possibly leading to more effective treatment strategies.
4 Conclusion
This investigation has demonstrated the significant bioactive potential of marine A. niger AMG31 isolated from Red Sea sediments. The fungal extract revealed a rich metabolite profile dominated by hesperetin and rosmarinic acid, alongside substantial quantities of phenolics, flavonoids, and tannins. These compounds likely contribute to the extract’s remarkable biological activities.
The green synthesis of Copper–zinc oxide nanocomposites using this fungal extract represents a sustainable approach that yielded crystalline, spherical nanostructures with sizes ranging from 12 to 45 nm. The physicochemical characterization confirmed the successful formation of the CZ nanocomposite with excellent stability, as evidenced by the zeta potential value of −23.5 mV. Notably, the exceptional hemocompatibility of the CZ nanocomposite (1.7% hemolysis at 1,000 μg·mL−1) underscores its potential suitability for biomedical applications. The fungal extract and CZ nanocomposite exhibited substantial antioxidant capacities. The anti-inflammatory assessment revealed that the CZ nanocomposite significantly outperformed the fungal extract in both COX-1 and COX-2 inhibition, suggesting its potential as an anti-inflammatory agent. The antimicrobial evaluation demonstrated that the fungal extract exhibited broader spectrum activity against most bacterial strains, while the CZ nanocomposite showed higher potency against specific pathogens, particularly E. faecalis and C. albicans.
These findings establish marine A. niger AMG31 as a valuable source of bioactive compounds with pharmaceutical potential and demonstrate the efficacy of fungal-mediated green synthesis of functional nanomaterials. The distinctive biological activities observed between the fungal extract and its derived nanocomposite highlight their potential applications in developing novel antimicrobial, antioxidant, and anti-inflammatory agents. Future research should focus on elucidating the molecular mechanisms underlying these biological activities and exploring their therapeutic applications through in vivo models and clinical studies.
Acknowledgments
This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) (grant number IMSIU‐DDRSP2501).
-
Funding information: This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) (grant number IMSIU‐DDRSP2501).
-
Author contributions: R.H.N.: methodology, investigation, and formal analysis. M.I.A.: methodology, investigation, and validation. Z.A.: methodology, visualization, and data curation. H.A.A.: investigation and methodology. N.A.T.: conceptualization and writing – review and editing. F.M.K.A.: data curation, formal analysis, and visualization. I.A.: data curation, formal analysis, and visualization. T.A.Y.: conceptualization, supervision, and project administration. I.M.I.: formal analysis and investigation. A.A.: methodology and data curation. A.F.: writing – review and editing. A.G.: conceptualization, methodology, investigation, supervision, writing – original draft, and writing – review and editing. All authors have accepted responsibility for the entire content of this manuscript and approved its submission.
-
Conflict of interest: The authors state no conflict of interest.
-
Data availability statement: The datasets generated and/or analyzed during the current study are available from the corresponding author on reasonable request.
References
[1] Wijayawardene, N. N., D.-Q. Dai, P. K. Jayasinghe, S. S. Gunasekara, Y. Nagano, S. Tibpromma, et al. Ecological and oceanographic perspectives in future marine fungal taxonomy. Journal of Fungi, Vol. 8, 2022, id. 1141.10.3390/jof8111141Suche in Google Scholar PubMed PubMed Central
[2] Mohamed, B. and N. Skliris. Recent sea level changes in the Red Sea: Thermosteric and halosteric contributions, and impacts of natural climate variability. Progress in Oceanography, Vol. 231, 2025, id. 103416.10.1016/j.pocean.2025.103416Suche in Google Scholar
[3] He, S., M. Ma, S. Epstein, and Y. Yin. Editorial: Microbes from marine distinctive environments. Frontiers in Microbiology, Vol. 14, 2023, id. 1300210.10.3389/fmicb.2023.1300210Suche in Google Scholar PubMed PubMed Central
[4] Kamat, S., S. Kumar, S. Philip, and M. Kumari. Chapter 10 - Secondary metabolites from marine fungi: current status and application. In: Microbial biomolecules, Kumar A., M. Bilal, L. F. R. Ferreira, and M. Kumari, eds, Elsevier, Amsterdam, Netherlands, 2023, pp. 181–209.10.1016/B978-0-323-99476-7.00001-6Suche in Google Scholar
[5] Dos Santos, I. R., T. A. Mohamed, L. L. Borges, H. A. Nahas, M. A. Abdel-Azeem, C. Carvalho, et al. The global tendency in the research of biological activity in endophytic fungi: a scientometric analysis. Current Research in Environmental and Applied Mycology, Vol. 12, 2022, pp. 1–14.10.5943/cream/12/1/1Suche in Google Scholar
[6] Wang, B., J. Cai, L. Huang, Y. Chen, R. Wang, M. Luo, et al. Significance of research on natural products from marine-derived Aspergillus species as a source against pathogenic bacteria. Frontiers in Microbiology, Vol. 15, 2024, id. 1464135.10.3389/fmicb.2024.1464135Suche in Google Scholar PubMed PubMed Central
[7] Abo Nahas, H. H., H. F. Abd El-Kareem, Y. H. Abo Nahas, M. Sherif, and A. M. Abdel-Azeem. Chapter 15 - Endophytic fungi: The budding source of natural antioxidants. In: Endophytic fungi, Abdel Azeem, A. M., A. N. Yadav, and N. Yadav, eds, Elsevier, Amsterdam, Netherlands, 2024, pp. 305–322.10.1016/B978-0-323-99314-2.00012-7Suche in Google Scholar
[8] Yu, R., J. Liu, Y. Wang, H. Wang, and H. Zhang. Aspergillus niger as a Secondary Metabolite Factory. Frontiers in Chemistry, Vol. 9, 2021, id. 701022.10.3389/fchem.2021.701022Suche in Google Scholar PubMed PubMed Central
[9] Abo Nahas, H., M. Abdel Rahman, V. Gupta, and A. Abdel-Azeem. Myco-antioxidants: insights into the natural metabolic treasure and their biological effects. Sydowia, Vol. 75, 2023, pp. 151–179.Suche in Google Scholar
[10] Yi, J., Y. Zhang, X. Wang, C. Wang, C. Sun, J. Wu, et al. Isolation and biological activities of chemical constituents from Aspergillus niger. Phytochemistry Letters, Vol. 60, 2024, pp. 159–162.10.1016/j.phytol.2024.02.002Suche in Google Scholar
[11] Abady, M. M., D. M. Mohammed, T. N. Soliman, R. A. Shalaby, and F. A. Sakr. Sustainable synthesis of nanomaterials using different renewable sources. Bulletin of the National Research Centre, Vol. 49, 2025, id. 24.10.1186/s42269-025-01316-4Suche in Google Scholar
[12] Rezghi Rami, M., M. Meskini, B., and Ebadi Sharafabad. Fungal-mediated nanoparticles for industrial applications: synthesis and mechanism of action. Journal of Infection and Public Health, Vol. 17, 2024, id. 102536.10.1016/j.jiph.2024.102536Suche in Google Scholar PubMed
[13] Madhavi, A., M. Srinivasulu, P. C. Shankar, and V. Rangaswamy. Synthesis and applications of fungal-mediated nanoparticles. In: Microbial processes for synthesizing nanomaterials, Maddela, N. R., J. M. Rodríguez Díaz, M. C. Branco da Silva Montenegro, R. Prasad, eds, Springer Nature, Singapore, 2023, pp. 113–131.10.1007/978-981-99-2808-8_5Suche in Google Scholar
[14] Manjula, N. G., Tajunnisa, V. Mamani, C. A. Meghana, S. B. Mayegowda. Fungal-based synthesis to generate nanoparticles for nanobioremediation. In: Green nanoremediation: Sustainable management of environmental pollution, Policarpo Tonelli, F. M., A. Roy, H. C. Ananda Murthy, eds, Springer International Publishing, Cham, 2023, pp. 83–108.10.1007/978-3-031-30558-0_4Suche in Google Scholar
[15] Montasser, K., M. Alnoman, H. Abozaid, F. Kouadri, M. Kamal, H. Gomaa, et al. Biofilm formation in OXA-producing Acinetobacter baumannii isolates and their susceptibility to different antibiotics. Sylwan, Vol. 163, 2019, id. 18.Suche in Google Scholar
[16] Magdy, O. S., I. M. Moussa, H. A. Hussein, M. D. El-Hariri, A. Ghareeb, H. A. Hemeg, et al. Genetic diversity of Salmonella enterica recovered from chickens farms and its potential transmission to human. Journal of Infection and Public Health, Vol. 13, 2020, pp. 571–576.10.1016/j.jiph.2019.09.007Suche in Google Scholar PubMed
[17] Abass, S., R. Parveen, and S. Ahmad. Green-synthesized nanoparticles: Characterization and antifungal mechanism of action. In: Advances in antifungal drug development: Natural products with antifungal potential, Manzoor, N., ed, Springer Nature, Singapore, 2024, pp. 373–388.10.1007/978-981-97-5165-5_13Suche in Google Scholar
[18] Zhu, X., Y. Chen, D. Yu, W. Fang, W. Liao, and W. Pan. Progress in the application of nanoparticles for the treatment of fungal infections: A review. Mycology, Vol. 15, 2024, pp. 1–16.10.1080/21501203.2023.2285764Suche in Google Scholar PubMed PubMed Central
[19] Ngwenya, S., N. J. Sithole, K. Ramachela, D. M. N. Mthiyane, M. Mwanza, M. Singh, et al. Eco-friendly synthesis of ZnO, CuO, and ZnO/CuO nanoparticles using extract of spent Pleurotus ostreatus substrate, and their antioxidant and anticancer activities. Discover Nano, Vol. 20, 2025, id. 35.10.1186/s11671-025-04199-6Suche in Google Scholar PubMed PubMed Central
[20] Takele, E., R. Bogale, G. Shumi, and G. Rundasa. Green synthesis, characterization, and antibacterial activity of CuO/ZnO nanocomposite using Zingiber officinale Rhizome extract. Journal of Chemistry, Vol. 2023, 2023, pp. 1–15.10.1155/2023/3481389Suche in Google Scholar
[21] Cao, Y., H. A. Dhahad, M. A. El-Shorbagy, H. Q. Alijani, M. Zakeri, A. Heydari, et al. Green synthesis of bimetallic ZnO-CuO nanoparticles and their cytotoxicity properties. Scientific Reports, Vol. 11, 2021, id. 23479.10.1038/s41598-021-02937-1Suche in Google Scholar PubMed PubMed Central
[22] Chenxi, Z., A. Hemmat, N. Thi, and M. Afrand. Nanoparticle-enhanced drug delivery systems: An up-to-date review. Journal of Molecular Liquids, Vol. 424, 2025, id. 126999.10.1016/j.molliq.2025.126999Suche in Google Scholar
[23] Abdelsalam, S. I. and M. M. Bhatti. Synergistic progression of nanoparticle dynamics in stenosed arteries. Qualitative Theory of Dynamical Systems, Vol. 24, 2024, id. 6.10.1007/s12346-024-01147-0Suche in Google Scholar
[24] El Kot, M. A., A. M. Alsharif, Y. Abd Elmaboud, and S. I. Abdelsalam. Harnessing electroosmotic hybrid nanofluid dynamics in curved arteries: insights into biomedical flow enhancement. Frontiers in Nanotechnology, Vol. 6, 2024.10.3389/fnano.2024.1520183Suche in Google Scholar
[25] Chen, W., C. Gao, B. Zhou, and Y. He. Biogenic copper/zinc oxide nanocomposites from Bixa orellana: Anticancer effects through ROS generation and apoptosis induction in cervical carcinoma. Journal of Cluster Science, Vol. 36, 2025, id. 52.10.1007/s10876-025-02766-6Suche in Google Scholar
[26] Fathi, M., N. Elsheshtawy, S. Youssef, and A. Ghareeb. Multiplex PCR identification of eight clinically relevant Candida species isolated from groin Fungal Infections among Egyptian patients. IOSR Journal of Pharmacy and Biological Sciences, Vol. 11, 2016, pp. 9–16.10.9790/3008-1104040916Suche in Google Scholar
[27] Swathi, J., K. M. Sowjanya, K. Narendra, K. V. N. R. Reddy, and A. Krishna Satya. Isolation, identification & production of bioactive metabolites from marine fungi collected from coastal area of Andhra Pradesh, India. Journal of Pharmacy Research, Vol. `6, 2013, pp. 663–666.10.1016/j.jopr.2013.04.052Suche in Google Scholar
[28] Alharbi, M. A., A. A. Alrehaili, M. O. I. Albureikan, A. F. Gharib, H. Daghistani, M. M. Bakhuraysah, et al. In vitro studies on the pharmacological potential, anti-tumor, antimicrobial, and acetylcholinesterase inhibitory activity of marine-derived Bacillus velezensis AG6 exopolysaccharide. RSC Advances, Vol. 13, 2023, pp. 26406–26417.10.1039/D3RA04009GSuche in Google Scholar
[29] Vishwakarma, S. Principle, instrumentation, and applications of gas chromatography-mass spectrometry (GC-MS), Vishakhapatnam University, India, 2021.Suche in Google Scholar
[30] Alshawwa, S. Z., K. S. Alshallash, A. Ghareeb, A. M. Elazzazy, M. Sharaf, A. Alharthi, et al. Assessment of pharmacological potential of novel exopolysaccharide isolated from Marine Kocuria sp. Strain AG5: Broad-spectrum biological investigations. Life, Vol. 12, 2022, id. 1387.10.3390/life12091387Suche in Google Scholar PubMed PubMed Central
[31] Sembiring, E., B. Elya, R. Sauriasari, and R. Sauriasari. Phytochemical screening, total flavonoid and total phenolic content and antioxidant activity of different parts of Caesalpinia bonduc (L.) Roxb. Pharmacognosy Journal, Vol. 10, 2018, pp. 123–127.10.5530/pj.2018.1.22Suche in Google Scholar
[32] Rover, M. and R. Brown. Quantification of total phenols in bio-oil using the Folin–Ciocalteu method. Journal of Analytical and Applied Pyrolysis, Vol. 104, 2013, pp. 366–371.10.1016/j.jaap.2013.06.011Suche in Google Scholar
[33] Chavan, U., F. Shahidi, and M. Naczk. Extraction of condensed tannins from beach pea (Lathyrus maritimus L.) as affected by different solvents. Food Chemistry, Vol. 75, 2001, pp. 509–512.10.1016/S0308-8146(01)00234-5Suche in Google Scholar
[34] Le, A., S. Parks, M. Nguyen, and P. Roach. Improving the vanillin-sulphuric acid method for quantifying total saponins. Technologies, Vol. 6, 2018, id. 84.10.3390/technologies6030084Suche in Google Scholar
[35] Ajanal, M., M. Gundkalle, and S. Nayak. Estimation of total alkaloid in Chitrakadivati by UV-Spectrophotometer. Ancient Science of Life, Vol. 31, 2012, pp. 198–201.10.4103/0257-7941.107361Suche in Google Scholar PubMed PubMed Central
[36] Šebesta, M., H. Vojtková, V. Cyprichová, A. P. Ingle, M. Urík, and M. Kolenčík. Mycosynthesis of metal-containing nanoparticles—fungal metal resistance and mechanisms of synthesis. International Journal of Molecular Sciences, Vol. 23, 2022, id. 14084.10.3390/ijms232214084Suche in Google Scholar PubMed PubMed Central
[37] Alharbi, M. A., A. M. A. Shahlol, M. O. I. Albureikan, K. Johani, M. G. Elsehrawy, M. El-Nablaway, et al. Investigating the multi-targeted pharmacological profile of an exopolysaccharide from Bacillus rugosus SYG20 via in vitro evaluation of its antioxidant, anti-inflammatory, anti-diabetic, wound healing, and antimicrobial properties. Archives of Medical Science, Vol. 21, No. 6, 2025.10.5114/aoms/190065Suche in Google Scholar
[38] Fouda, A., E. Saied, A. M. Eid, F. Kouadri, A. M. Alemam, M. F. Hamza, et al. Green synthesis of zinc oxide nanoparticles using an aqueous extract of Punica granatum for antimicrobial and catalytic activity. Journal of Functional Biomaterials, Vol. 14, 2023, id. 205.10.3390/jfb14040205Suche in Google Scholar PubMed PubMed Central
[39] Kumari, N., V. Sudharsan, T. Muthu Kutty, N. Jayan, and M. Laxmi Deepak Bhatlu. Green synthesis and characterization of Zinc and Copper oxides nanocomposite using Phyllanthus emblica extracts and its antibacterial and antioxidant properties. Materials Today: Proceedings, 2023.10.1016/j.matpr.2023.06.287Suche in Google Scholar
[40] Yin, X., A. M. Eid, Y. Wei, M. F. Hamza, M. A. Abdel-Rahman, N. Zheng, et al. Biogenic iron oxide nanoparticles with enhanced photocatalytic activity for effective removal of uranium and microbial decontamination. Journal of Water Process Engineering, Vol. 70, 2025, id. 107093.10.1016/j.jwpe.2025.107093Suche in Google Scholar
[41] Amin, F., B. Fozia, A. Khattak, M. Alotaibi, I. Qasim, R. Ahmad, et al. Green synthesis of copper oxide nanoparticles using Aerva javanica leaf extract and their characterization and investigation of in vitro antimicrobial potential and cytotoxic activities. Evidence-based Complementary and Alternative Medicine : eCAM, Vol. 2021, 2021, id. 5589703.10.1155/2021/5589703Suche in Google Scholar PubMed PubMed Central
[42] Ghareeb, A., A. Fouda, R. M. Kishk, and W. M. El Kazzaz. Multifaceted biomedical applications of biogenic titanium dioxide nanoparticles fabricated by marine actinobacterium Streptomyces vinaceusdrappus AMG31. Scientific Reports, Vol. 15, 2025, id. 20244.10.1038/s41598-025-00541-1Suche in Google Scholar PubMed PubMed Central
[43] Khowdiary, M. M., Z. Alatawi, A. Alhowiti, M. A. Amin, H. Daghistani, F. M. K. Albaqami, et al. Phytochemical analysis and multifaceted biomedical activities of Nitraria retusa extract as natural product-based therapies. Life, Vol. 14, 2024, id. 1629.10.3390/life14121629Suche in Google Scholar PubMed PubMed Central
[44] Alharbi, N. K., Z. F. Azeez, H. M. Alhussain, A. M. A. Shahlol, M. O. I. Albureikan, M. G. Elsehrawy, et al. Tapping the biosynthetic potential of marine Bacillus licheniformis LHG166, a prolific sulphated exopolysaccharide producer: structural insights, bio-prospecting its antioxidant, antifungal, antibacterial and anti-biofilm potency as a novel anti-infective lead. Frontiers in Microbiology, Vol. 15, 2024, id. 1385493.10.3389/fmicb.2024.1385493Suche in Google Scholar PubMed PubMed Central
[45] Aloraini, G. S., M. O. I. Albureikan, A. M. A. Shahlol, T. Shamrani, H. Daghistani, M. El-Nablaway, et al. Biomedical and therapeutic potential of marine-derived Pseudomonas sp. strain AHG22 exopolysaccharide: A novel bioactive microbial metabolite. Reviews on Advanced Materials Science, Vol. 63, 2024, id. 20240016.10.1515/rams-2024-0016Suche in Google Scholar
[46] Hussein Naser, R., Z. Falah Azeez, Z. Alatawi, A. Albalawi, T. Shamrani, A. M. A. Shahlol, et al. Oxidative stress-induced DNA damage and apoptosis in multiple cancer cell lines: novel anticancer properties of marine Aspergillus oryzae NGM91 extract. RSC Advances, Vol. 15, 2025, pp. 17203–17221.10.1039/D5RA02028JSuche in Google Scholar
[47] Alwaili, M. A., M. A. Alshehri, T. R. Abdulrahman, F. M. K. Albaqami, A. Alghamdi, M. O. I. Albureikan, et al. Broad-spectrum bioactivities of a sulfated heteropolysaccharide from Bacillus tequilensis MYG163: antioxidant, anti-Inflammatory, anticancer, antimicrobial, and antibiofilm properties. Journal of Taibah University for Science, Vol. 19, 2025, id. 2447151.10.1080/16583655.2024.2447151Suche in Google Scholar
[48] Ghareeb, A., A. Fouda, R. M. Kishk, and W. M. El Kazzaz. Unlocking the therapeutic potential of bioactive exopolysaccharide produced by marine actinobacterium Streptomyces vinaceusdrappus AMG31: A novel approach to drug development. International Journal of Biological Macromolecules, Vol. 276, 2024, id. 133861.10.1016/j.ijbiomac.2024.133861Suche in Google Scholar PubMed
[49] Hamza, M. F., A. Fouda, Y. Wei, I. E. El Aassy, S. H. Alotaibi, E. Guibal, et al. Functionalized biobased composite for metal decontamination – Insight on uranium and application to water samples collected from wells in mining areas (Sinai, Egypt). Chemical Engineering Journal, Vol. 431, 2022, id. 133967.10.1016/j.cej.2021.133967Suche in Google Scholar
[50] Coates, J. Interpretation of infrared spectra, a practical approach. In: Encyclopedia of analytical chemistry, John Wiley & Sons, Inc., Hoboken, New Jersey, United States, 2006.Suche in Google Scholar
[51] Senthilkumaar, S., K. Rajendran, S. Banerjee, T. K. Chini, and V. Sengodan. Influence of Mn doping on the microstructure and optical property of ZnO. Materials Science in Semiconductor Processing, Vol. 11, 2008, pp. 6–12.10.1016/j.mssp.2008.04.005Suche in Google Scholar
[52] Abdel-Wahab, B., H. Abd El-Kareem, A. Alzamami, C. Fahmy, B. Elesawy, M. Mahmoud, et al. Novel exopolysaccharide from Marine Bacillus subtilis with broad potential biological activities: Insights into antioxidant, anti-inflammatory, cytotoxicity, and anti-Alzheimer activity. Metabolites, Vol. 12, 2022, id. 715.10.3390/metabo12080715Suche in Google Scholar PubMed PubMed Central
[53] Selim, S., M. S. Almuhayawi, M. T. Alharbi, M. K. Nagshabandi, A. Alanazi, M. Warrad, et al. In vitro assessment of antistaphylococci, antitumor, immunological and structural characterization of acidic bioactive exopolysaccharides from Marine Bacillus cereus Isolated from Saudi Arabia. Metabolites, Vol. 12, 2022, id. 132.10.3390/metabo12020132Suche in Google Scholar PubMed PubMed Central
[54] Atiek, E., A. Matebu, D. Tsegaye, G. Behailu, and B. Abebe. Green synthesis of TiO2/ZnO heterostructure using Urtica Smensis leaf extract for antibacterial activity. Results in Chemistry, Vol. 12, 2024, id. 101880.10.1016/j.rechem.2024.101880Suche in Google Scholar
[55] Muthukumaran, S. and R. Gopalakrishnan. Structural, FTIR and photoluminescence studies of Cu doped ZnO nanopowders by co-precipitation method. Optical Materials (Amsterdam, Netherlands), Vol. 34, 2012, pp. 1946–1953.10.1016/j.optmat.2012.06.004Suche in Google Scholar
[56] Haghighizadeh, A., A. Aghababai Beni, M. Haghmohammadi, M. S. S. Adel, and S. Farshad. Green synthesis of ZnO-TiO2 nano-photocatalyst doped with Fe(III) ions using bitter olive extract to treat textile wastewater containing reactive dyes. Water, Air, and Soil Pollution, Vol. 234, 2023, id. 366.10.1007/s11270-023-06374-wSuche in Google Scholar
[57] Sonkar, R., N. J. Mondal, B. Boro, M. P. Ghosh, and D. Chowdhury. Cu doped ZnO nanoparticles: Correlations between tuneable optoelectronic, antioxidant and photocatalytic activities. Journal of Physics and Chemistry of Solids, Vol. 185, 2024, id. 111715.10.1016/j.jpcs.2023.111715Suche in Google Scholar
[58] Elavarthi, P., R. Mangiri, A. Sudharani, and S. Kummara. Nano synthesis and characterization of Co and Mn Co-doped ZnO by solution combustion technique. Journal of Superconductivity and Novel Magnetism, Vol. 34, 2021, pp. 342–352.10.1007/s10948-021-05874-2Suche in Google Scholar
[59] Pragathiswaran, C., C. Smitha, B. Mahin Abbubakkar, P. Govindhan, and N. Anantha Krishnan. Synthesis and characterization of TiO2/ZnO–Ag nanocomposite for photocatalytic degradation of dyes and anti-microbial activity. Materials Today: Proceedings, Vol. 45, 2021, pp. 3357–3364.10.1016/j.matpr.2020.12.664Suche in Google Scholar
[60] Ghareeb, A., A. Fouda, R. M. Kishk, and W. M. El Kazzaz. Unlocking the potential of titanium dioxide nanoparticles: an insight into green synthesis, optimizations, characterizations, and multifunctional applications. Microbial Cell Factories, Vol. 23, 2024, id. 341.10.1186/s12934-024-02609-5Suche in Google Scholar PubMed PubMed Central
[61] Mohd Yusof, H., R. Mohamad, U. H. Zaidan, and N. A. Abdul Rahman. Microbial synthesis of zinc oxide nanoparticles and their potential application as an antimicrobial agent and a feed supplement in animal industry: a review. Journal of Animal Science and Biotechnology, Vol. 10, 2019, id. 57.10.1186/s40104-019-0368-zSuche in Google Scholar PubMed PubMed Central
[62] Turabik, M., S. Özdemir, G. Akinbingol, S. Gonca, and C. Gecgel. Comparison of antioxidant, antimicrobial, DNA cleavage, cell viability, and biofilm inhibition activities of mono- and bimetallic copper and zinc nanoparticles. Inorganic Chemistry Communications, Vol. 155, 2023, id. 111072.10.1016/j.inoche.2023.111072Suche in Google Scholar
[63] Madeshwaran, K. and R. Venkatachalam. Green synthesis of bimetallic ZnO–CuO nanoparticles using Annona muricata l. extract: Investigation of antimicrobial, antioxidant, and anticancer properties. Journal of Industrial and Engineering Chemistry, Vol. 140, 2024, pp. 454–467.10.1016/j.jiec.2024.06.002Suche in Google Scholar
[64] Jeevarathinam, M. and I. V. Asharani. Synthesis of CuO, ZnO nanoparticles, and CuO-ZnO nanocomposite for enhanced photocatalytic degradation of Rhodamine B: a comparative study. Scientific Reports, Vol. 14, 2024, id. 9718.10.1038/s41598-024-60008-7Suche in Google Scholar PubMed PubMed Central
[65] Fouda, A., A. M. Eid, E. Guibal, M. F. Hamza, S. E.-D. Hassan, D. H. M. Alkhalifah, et al. Green synthesis of gold nanoparticles by aqueous extract of Zingiber officinale: characterization and insight into antimicrobial, antioxidant, and in vitro cytotoxic activities. Applied Sciences, Vol. 12, 2022, id. 12879.10.3390/app122412879Suche in Google Scholar
[66] Danaei, M., M. Dehghankhold, S. Ataei, F. Hasanzadeh Davarani, R. Javanmard, A. Dokhani, et al. Impact of particle size and polydispersity index on the clinical applications of lipidic nanocarrier systems. Pharmaceutics, Vol. 10, 2018, id. 57.10.3390/pharmaceutics10020057Suche in Google Scholar PubMed PubMed Central
[67] Bhattacharjee, S. DLS and zeta potential – What they are and what they are not? Journal of Controlled Release, Vol. 235, 2016, pp. 337–351.10.1016/j.jconrel.2016.06.017Suche in Google Scholar PubMed
[68] Ghaznavi, H., M. R. Hajinezhad, Z. Hesari, M. Shirvaliloo, S. Sargazi, S. Shahraki, et al. Synthesis, characterization, and evaluation of copper-doped zinc oxide nanoparticles anticancer effects: in vitro and in vivo experiments. BMC Cancer, Vol. 25, 2025, id. 37.10.1186/s12885-024-13398-wSuche in Google Scholar PubMed PubMed Central
[69] Mishra, D. N., L. Prasad, and U. Suyal. Synthesis of zinc oxide nanoparticles using Trichoderma harzianum and its bio-efficacy on Alternaria brassicae. Frontiers in Microbiology, Vol. 16, 2025, id. 1506695.10.3389/fmicb.2025.1506695Suche in Google Scholar PubMed PubMed Central
[70] Sykuła, A., A. Bodzioch, A. Nowak, W. Maniukiewicz, S. Ścieszka, L. Piekarska-Radzik, et al. Encapsulation and biological activity of hesperetin derivatives with HP-β-CD. Molecules, Vol. 28, 2023, id. 6893.10.3390/molecules28196893Suche in Google Scholar PubMed PubMed Central
[71] Kowalczyk, A., C. I. G. Tuberoso, and I. Jerković. The role of Rosmarinic acid in cancer prevention and therapy: mechanisms of antioxidant and anticancer activity. Antioxidants, Vol. 13, 2024, id. 1313.10.3390/antiox13111313Suche in Google Scholar PubMed PubMed Central
[72] Sirikhansaeng, P., T. Tanee, R. Sudmoon, and A. Chaveerach. Major phytochemical as γ-sitosterol disclosing and toxicity testing in Lagerstroemia species. Evidence-based Complementary and Alternative Medicine : eCAM, Vol. 2017, 2017, id. 7209851.10.1155/2017/7209851Suche in Google Scholar PubMed PubMed Central
[73] Khameneh, B., N. A. M. Eskin, M. Iranshahy, and B. S. Fazly Bazzaz. Phytochemicals: A promising weapon in the arsenal against antibiotic-resistant bacteria. Antibiotics (USSR), Vol. 10, 2021, id. 1044.10.3390/antibiotics10091044Suche in Google Scholar PubMed PubMed Central
[74] Ajijolakewu, A. K., M. O. Kazeem, G. M. Dovia, I. A. Adebayo, and N. T. Ajide-Bamigboye. Structural characterization and evaluation of mycogenic zinc oxide nanoparticles from the cell-free culture-extract of Aspergillus niger. Ife Journal of Science, Vol. 26, 2024, pp. 131–143.10.4314/ijs.v26i1.9Suche in Google Scholar
[75] Yan, H., W. Hou, B. Lei, J. Liu, R. Song, W. Hao, et al. Ultrarobust stable ABTS radical cation prepared using Spore@Cu-TMA biocomposites for antioxidant capacity assay. Talanta, Vol. 276, 2024, id. 126282.10.1016/j.talanta.2024.126282Suche in Google Scholar PubMed
[76] Kavipriya, M., D. Selleswari, U. Unnimaya, and R. K. S. J. R. Antioxidant, antibacterial and anti-cancer properties of bio -fabricated copper-silver-zinc nanocomposites: An ecofriendly approach. Uttar Pradesh Journal of Zoology, Vol. 45, 2024, pp. 15–27.10.56557/upjoz/2024/v45i164283Suche in Google Scholar
[77] Suresha, B. S. and K. Srinivasan. Antioxidant properties of fungal metabolite nigerloxin in vitro. Prikladnaia biokhimiia i mikrobiologiia, Vol. 49, 2013, pp. 587–591.10.1134/S0003683813060173Suche in Google Scholar PubMed
[78] Chakraborty, A., S. Majumdar, and J. Bhowal. Phytochemical screening and antioxidant and antimicrobial activities of crude extracts of different filamentous fungi. Archives of Microbiology, Vol. 203, 2021, pp. 6091–6108.10.1007/s00203-021-02572-4Suche in Google Scholar PubMed
[79] Yang, T., K. Yang, Y. Zhang, R. Zhou, F. Zhang, G. Zhan, et al. Metabolites with antioxidant and α-glucosidase inhibitory activities produced by the endophytic fungi Aspergillus niger from Pachysandra terminalis. Bioscience, Biotechnology, and Biochemistry, Vol. 86, 2022, pp. 1343–1348.10.1093/bbb/zbac137Suche in Google Scholar PubMed
[80] Gomez Rojas, M. P. and O. Martinez. Whey protein fermentation with Aspergillus niger: Source of antioxidant peptides, IntechOpen Limited, London, United Kingdom, 2023.Suche in Google Scholar
[81] Pandey, P., S. Lakhanpal, D. Mahmood, H. N. Kang, B. Kim, S. Kang, et al. An updated review summarizing the anticancer potential of flavonoids via targeting NF-kB pathway. Frontiers in Pharmacology, Vol. 15, 2025, id. 1513422.10.3389/fphar.2024.1513422Suche in Google Scholar PubMed PubMed Central
[82] Khan, S., K. L. Andrews, and J. P. F. Chin-Dusting. Cyclo-oxygenase (COX) inhibitors and cardiovascular risk: are non-steroidal anti-inflammatory drugs really anti-inflammatory? International Journal of Molecular Sciences, Vol. 20, 2019, id. 4262.10.3390/ijms20174262Suche in Google Scholar PubMed PubMed Central
[83] Adeyemi, J. O., D. C. Onwudiwe, and A. O. Oyedeji. Biogenic synthesis of CuO, ZnO, and CuO-ZnO nanoparticles using leaf extracts of Dovyalis caffra and their biological properties. Molecules, Vol. 27, 2022, id. 3206.10.3390/molecules27103206Suche in Google Scholar PubMed PubMed Central
[84] Krishnasamy, L., S. Sundaram, and B. Ravichandran. Biogenic copper nanoparticles from Combined leaf extracts of Catharanthus roseus (L.) and Abutilon Indicum(L.) and their biological effects -an In vitro study. Tuijin Jishu/Journal of Propulsion Technology, Vol. 44, 2023, pp. 1001–4055.10.52783/tjjpt.v44.i4.2282Suche in Google Scholar
[85] Duran, B., E. Perçin, A. Aydın, A. Destegül, D. Aydin, N. Özpınar, et al. Biogenic nanocopper: Eco-friendly synthesis, characterization, and their anti-trichomonas vaginalis, anticancer, and antimicrobial effects. Journal of Medicinal Food, Vol. 25, 2022, pp. 787–792.10.1089/jmf.2021.0151Suche in Google Scholar PubMed
[86] Abou Zaid, E. S., S. Z. Mansour, S. M. El-Sonbaty, F. S. Moawed, E. I. Kandil, and R. A.-H. Haroun. Boswellic acid coated zinc nanoparticles attenuate NF-κB-mediated inflammation in DSS-induced ulcerative colitis in rats. International Journal of Immunopathology and Pharmacology, Vol. 37, 2023, id. 3946320221150720.10.1177/03946320221150720Suche in Google Scholar PubMed PubMed Central
[87] Chang, C. J. and D. C. Brady. Capturing copper to inhibit inflammation. Nature Chemical Biology, Vol. 19, 2023, pp. 937–939.10.1038/s41589-023-01383-6Suche in Google Scholar PubMed
[88] Hussain, A., M. F. AlAjmi, M. T. Rehman, S. Amir, F. M. Husain, A. Alsalme, et al. Copper(II) complexes as potential anticancer and Nonsteroidal anti-inflammatory agents: In vitro and in vivo studies. Scientific Reports, Vol. 9, 2019, id. 5237.10.1038/s41598-019-41063-xSuche in Google Scholar PubMed PubMed Central
[89] Mohapatra, A., S. K. Rajendrakumar, G. Chandrasekaran, V. Revuri, P. Sathiyamoorthy, Y.-K. Lee, et al. Biomineralized nanoscavenger abrogates proinflammatory macrophage polarization and induces neutrophil clearance through reverse migration during gouty arthritis. ACS Applied Materials & Interfaces, Vol. 15, 2023, pp. 3812–3825.10.1021/acsami.2c19684Suche in Google Scholar PubMed
[90] Bonaccorsi, M., M. J. Knight, T. Le Marchand, H. R. W. Dannatt, T. Schubeis, L. Salmon, et al. Multimodal response to copper binding in superoxide dismutase dynamics. Journal of the American Chemical Society, Vol. 142, 2020, pp. 19660–19667.10.1021/jacs.0c09242Suche in Google Scholar PubMed
[91] Ghareeb, A., M. Sayed, E. Sayed, and K. Montasser. Impact of human papilloma virus HPV on immunosenescent CD57 + T-lymphocytes in cervical cancer patients. Microbial Biosystems, Vol. 6, 2021, pp. 11–21.10.21608/mb.2021.106010.1046Suche in Google Scholar
[92] Liu, Y., J. He, M. Li, K. Ren, and Z. Zhao. Inflammation-driven nanohitchhiker enhances postoperative immunotherapy by alleviating prostaglandin E2-mediated immunosuppression. ACS Applied Materials & Interfaces, Vol. 16, 2024, pp. 6879–6893.10.1021/acsami.3c17357Suche in Google Scholar PubMed
[93] Souffriau, J., S. Timmermans, T. Vanderhaeghen, C. Wallaeys, K. Van Looveren, L. Aelbrecht, et al. Zinc inhibits lethal inflammatory shock by preventing microbe-induced interferon signature in intestinal epithelium. EMBO Molecular Medicine, Vol. 12, 2020, id. e11917.10.15252/emmm.201911917Suche in Google Scholar PubMed PubMed Central
[94] Gamal, M., G. Al-Sayyad, A. Elmetwalli, S. A. Abo-ElMaaty, and M. G. Hassan. Copper-zinc nanoparticles: Synthesis, physicochemical properties, and biological efficacy against bacteria and cancer cells. Journal of Basic and Environmental Sciences, Vol. 11, 2024, pp. 813–826.10.21608/jbes.2024.396165Suche in Google Scholar
[95] de, B., M. Santana, G. M. Armentano, D. A. S. Ferreira, C. S. de Freitas, M. S. Carneiro-Ramos, et al. In vitro bactericidal activity of biogenic copper oxide nanoparticles for Neisseria gonorrhoeae with enhanced compatibility for human cells. ACS Applied Materials & Interfaces, Vol. 16, 2024, pp. 21633–21642.10.1021/acsami.4c02357Suche in Google Scholar PubMed
[96] Rasheed, R., A. Bhat, B. Singh, and F. Tian. Biogenic synthesis of selenium and copper oxide nanoparticles and inhibitory effect against multi-drug resistant biofilm-forming bacterial pathogens. Biomedicines, Vol. 12, 2024, id. 994.10.3390/biomedicines12050994Suche in Google Scholar PubMed PubMed Central
[97] Biswas, R. Antimicrobial activity of biosynthesized copper nanoparticles using methanolic extract of Ocimum Sanctum. International Journal for Research in Applied Science and Engineering Technology, Vol. 12, 2024, pp. 466–472.10.22214/ijraset.2024.63600Suche in Google Scholar
[98] Nabila, M. I. and K. Kannabiran. Biosynthesis, characterization and antibacterial activity of copper oxide nanoparticles (CuO NPs) from actinomycetes. Biocatalysis and Agricultural Biotechnology, Vol. 15, 2018, pp. 56–62.10.1016/j.bcab.2018.05.011Suche in Google Scholar
[99] Garcia-Marin, L. E., K. Juarez-Moreno, A. R. Vilchis-Nestor, and E. Castro-Longoria. Highly antifungal activity of biosynthesized copper oxide nanoparticles against Candida albicans. Nanomaterials, Vol. 12, 2022, id. 3856.10.3390/nano12213856Suche in Google Scholar PubMed PubMed Central
[100] Yassin, M. T., F. O. Al-Otibi, A. A. Al-Askar, and M. M. Elmaghrabi. Synergistic anticandidal effectiveness of greenly synthesized zinc oxide nanoparticles with antifungal agents against nosocomial candidal pathogens. Microorganisms, Vol. 11, 2023, id. 1957.10.3390/microorganisms11081957Suche in Google Scholar PubMed PubMed Central
[101] Rojas-Jaimes, J. and D. Asmat-Campos. Cu2O, ZnO, and Ag/Cu2O nanoparticles synthesized by biogenic and chemical route and their effect on Pseudomonas aeruginosa and Candida albicans. Scientific Reports, Vol. 13, 2023, id. 21478.10.1038/s41598-023-47917-9Suche in Google Scholar PubMed PubMed Central
[102] Shehabeldine, A. M., B. H. Amin, F. A. Hagras, A. A. Ramadan, M. R. Kamel, M. A. Ahmed, et al. Potential antimicrobial and antibiofilm properties of copper oxide nanoparticles: Time-kill kinetic essay and ultrastructure of pathogenic bacterial cells. Applied Biochemistry and Biotechnology, Vol. 195, 2023, pp. 467–485.10.1007/s12010-022-04120-2Suche in Google Scholar PubMed PubMed Central
[103] Metryka, O., D. Wasilkowski, M. Dulski, M. Adamczyk-Habrajska, M. Augustyniak, and A. Mrozik. Metallic nanoparticle actions on the outer layer structure and properties of Bacillus cereus and Staphylococcus epidermidis. Chemosphere, Vol. 354, 2024, id. 141691.10.1016/j.chemosphere.2024.141691Suche in Google Scholar PubMed
[104] Koçak, Y. and H. Seçkin. Synthesis, characterization of biogenic copper nanoparticles and their therapeutic activity. Journal of Agriculture, Vol. 6, 2023, pp. 26–35.10.46876/ja.1287833Suche in Google Scholar
[105] Sharma, A., K. Akash, S. Kumari, K. Chauhan, A. James, R. Goel, et al. Biogenic zinc oxide nanoparticles: an insight into the advancements in antimicrobial resistance. ECS Journal of Solid State Science and Technology, Vol. 13, 2024, id. 047002.10.1149/2162-8777/ad397fSuche in Google Scholar
[106] Qu, X., H. Yang, B. Jia, Z. Yu, Y. Zheng, and K. Dai. Biodegradable Zn–Cu alloys show antibacterial activity against MRSA bone infection by inhibiting pathogen adhesion and biofilm formation. Acta Biomaterialia, Vol. 117, 2020, pp. 400–417.10.1016/j.actbio.2020.09.041Suche in Google Scholar PubMed
[107] Summer, M., S. Ali, H. M. Tahir, R. Abaidullah, U. Fiaz, S. Mumtaz, et al. Mode of action of biogenic silver, zinc, copper, titanium and cobalt nanoparticles against antibiotics resistant pathogens. Journal of Inorganic and Organometallic Polymers and Materials, Vol. 34, 2024, pp. 1417–1451.10.1007/s10904-023-02935-ySuche in Google Scholar
[108] Zulqarnain, Z. Iqbal, R. Cox, J. Anwar, N. Ahmad, K. Khan, et al. Antifungal activity of compounds isolated from Aspergillus niger and their molecular docking studies with tomatinase. Natural Product Research, Vol. 34, 2020, pp. 2642–2646.10.1080/14786419.2018.1548447Suche in Google Scholar PubMed
[109] Alhasan, D., H. A. Hussein, and A. Dawood. In vitro antimicrobial activity of the filtrate crude extract produced by Aspergillus niger. University of Thi-Qar Journal of Science, Vol. 7, 2019, pp. 66–71.10.32792/utq/utjsci/v7i1.253Suche in Google Scholar
[110] Üstün, A., A. Yazici, and S. Örtucu. Isolation of a novel antimicrobial polypeptide from Aspergillus niger isolate. Trakya University Journal of Natural Sciences, Vol. 24, 2023, pp. 41–48.10.23902/trkjnat.1247186Suche in Google Scholar
[111] Carević, T., M. Kostić, B. Nikolić, D. Stojković, M. Soković, and M. Ivanov. Hesperetin-between the ability to diminish mono- and polymicrobial biofilms and toxicity. Molecules, Vol. 27, 2022, id. 6806.10.3390/molecules27206806Suche in Google Scholar PubMed PubMed Central
[112] Chen, X., W. Lan, and J. Xie. Natural phenolic compounds: Antimicrobial properties, antimicrobial mechanisms, and potential utilization in the preservation of aquatic products. Food Chemistry, Vol. 440, 2024, id. 138198.10.1016/j.foodchem.2023.138198Suche in Google Scholar PubMed
[113] Harrison, M. A., R. J. Farthing, N. Allen, L. M. Ahern, K. Birchall, M. Bond, et al. Identification of novel p-cresol inhibitors that reduce Clostridioides difficile’s ability to compete with species of the gut microbiome. Scientific Reports, Vol. 13, 2023, id. 9492.10.1038/s41598-023-32656-8Suche in Google Scholar PubMed PubMed Central
[114] Alshehri, M. M., J. Sharifi-Rad, J. Herrera-Bravo, E. L. Jara, L. A. Salazar, D. Kregiel, et al. Therapeutic potential of isoflavones with an emphasis on daidzein. Oxidative Medicine and Cellular Longevity, Vol. 2021, 2021, id. 6331630.10.1155/2021/6331630Suche in Google Scholar PubMed PubMed Central
[115] Xie, Y. Daidzin: Advances on resources, biosynthesis pathway, bioavailability, bioactivity, and pharmacology. In: Handbook of dietary flavonoids, Xiao, J, ed, Springer International Publishing, Cham, 2023, pp. 1–22.10.1007/978-3-030-94753-8_36-1Suche in Google Scholar
[116] Bais, H., T. Walker, H. Schweizer, and J. Vivanco. Root specific elicitation and antimicrobial activity of rosmarinic acid in hairy root cultures of Ocimum basilicum. Plant Physiology and Biochemistry, Vol. 40, 2002, pp. 983–995.10.1016/S0981-9428(02)01460-2Suche in Google Scholar
[117] Singh, V. K., A. Mishra, and B. Jha. 3-Benzyl-hexahydro-pyrrolo[1,2-a] pyrazine-1,4-dione extracted From Exiguobacterium indicum showed anti-biofilm activity against Pseudomonas aeruginosa by attenuating quorum sensing. Frontiers in Microbiology, Vol. 10, 2019, id. 1269.10.3389/fmicb.2019.01269Suche in Google Scholar PubMed PubMed Central
[118] Balamurugan, R., A. Stalin, A. Aravinthan, and J.-H. Kim. γ-sitosterol a potent hypolipidemic agent: In silico docking analysis. Medicinal Chemistry Research, Vol. 24, 2015, pp. 124–130.10.1007/s00044-014-1075-0Suche in Google Scholar
[119] Limwongyut, J., A. S. Moreland, C. Nie, J. Read de Alaniz, and G. C. Bazan. Amide moieties modulate the antimicrobial activities of conjugated oligoelectrolytes against gram-negative bacteria. ChemistryOpen, Vol. 11, 2022, id. e202100260.10.1002/open.202100260Suche in Google Scholar PubMed PubMed Central
[120] de Melo, L. F., V. G. de Queiroz Aquino-Martins, A. P. da Silva, H. A. Oliveira Rocha, and K. C. Scortecci. Biological and pharmacological aspects of tannins and potential biotechnological applications. Food Chemistry, Vol. 414, 2023, id. 135645.10.1016/j.foodchem.2023.135645Suche in Google Scholar PubMed
[121] Farha, A. K., Q.-Q. Yang, G. Kim, H.-B. Li, F. Zhu, H.-Y. Liu, et al. Tannins as an alternative to antibiotics. Food Bioscience, Vol. 38, 2020, id. 100751.10.1016/j.fbio.2020.100751Suche in Google Scholar
© 2025 the author(s), published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Artikel in diesem Heft
- Review Articles
- Utilization of steel slag in concrete: A review on durability and microstructure analysis
- Technical development of modified emulsion asphalt: A review on the preparation, performance, and applications
- Recent developments in ultrasonic welding of similar and dissimilar joints of carbon fiber reinforcement thermoplastics with and without interlayer: A state-of-the-art review
- Unveiling the crucial factors and coating mitigation of solid particle erosion in steam turbine blade failures: A review
- From magnesium oxide, magnesium oxide concrete to magnesium oxide concrete dams
- Properties and potential applications of polymer composites containing secondary fillers
- A scientometric review on the utilization of copper slag as a substitute constituent of ordinary Portland cement concrete
- Advancement of additive manufacturing technology in the development of personalized in vivo and in vitro prosthetic implants
- Recent advance of MOFs in Fenton-like reaction
- A review of defect formation, detection, and effect on mechanical properties of three-dimensional braided composites
- Non-conventional approaches to producing biochars for environmental and energy applications
- Review of the development and application of aluminum alloys in the nuclear industry
- Advances in the development and characterization of combustible cartridge cases and propellants: Preparation, performance, and future prospects
- Recent trends in rubberized and non-rubberized ultra-high performance geopolymer concrete for sustainable construction: A review
- Cement-based materials for radiative cooling: Potential, material and structural design, and future prospects
- A comprehensive review: The impact of recycling polypropylene fiber on lightweight concrete performance
- A comprehensive review of preheating temperature effects on reclaimed asphalt pavement in the hot center plant recycling
- Research Articles
- Investigation of the corrosion performance of HVOF-sprayed WC-CoCr coatings applied on offshore hydraulic equipment
- A systematic review of metakaolin-based alkali-activated and geopolymer concrete: A step toward green concrete
- Evaluation of color matching of three single-shade composites employing simulated 3D printed cavities with different thicknesses using CIELAB and CIEDE2000 color difference formulae
- Novel approaches in prediction of tensile strain capacity of engineered cementitious composites using interpretable approaches
- Effect of TiB2 particles on the compressive, hardness, and water absorption responses of Kulkual fiber-reinforced epoxy composites
- Analyzing the compressive strength, eco-strength, and cost–strength ratio of agro-waste-derived concrete using advanced machine learning methods
- Tensile behavior evaluation of two-stage concrete using an innovative model optimization approach
- Tailoring the mechanical and degradation properties of 3DP PLA/PCL scaffolds for biomedical applications
- Optimizing compressive strength prediction in glass powder-modified concrete: A comprehensive study on silicon dioxide and calcium oxide influence across varied sample dimensions and strength ranges
- Experimental study on solid particle erosion of protective aircraft coatings at different impact angles
- Compatibility between polyurea resin modifier and asphalt binder based on segregation and rheological parameters
- Fe-containing nominal wollastonite (CaSiO3)–Na2O glass-ceramic: Characterization and biocompatibility
- Relevance of pore network connectivity in tannin-derived carbons for rapid detection of BTEX traces in indoor air
- A life cycle and environmental impact analysis of sustainable concrete incorporating date palm ash and eggshell powder as supplementary cementitious materials
- Eco-friendly utilisation of agricultural waste: Assessing mixture performance and physical properties of asphalt modified with peanut husk ash using response surface methodology
- Benefits and limitations of N2 addition with Ar as shielding gas on microstructure change and their effect on hardness and corrosion resistance of duplex stainless steel weldments
- Effect of selective laser sintering processing parameters on the mechanical properties of peanut shell powder/polyether sulfone composite
- Impact and mechanism of improving the UV aging resistance performance of modified asphalt binder
- AI-based prediction for the strength, cost, and sustainability of eggshell and date palm ash-blended concrete
- Investigating the sulfonated ZnO–PVA membrane for improved MFC performance
- Strontium coupling with sulphur in mouse bone apatites
- Transforming waste into value: Advancing sustainable construction materials with treated plastic waste and foundry sand in lightweight foamed concrete for a greener future
- Evaluating the use of recycled sawdust in porous foam mortar for improved performance
- Improvement and predictive modeling of the mechanical performance of waste fire clay blended concrete
- Polyvinyl alcohol/alginate/gelatin hydrogel-based CaSiO3 designed for accelerating wound healing
- Research on assembly stress and deformation of thin-walled composite material power cabin fairings
- Effect of volcanic pumice powder on the properties of fiber-reinforced cement mortars in aggressive environments
- Analyzing the compressive performance of lightweight foamcrete and parameter interdependencies using machine intelligence strategies
- Selected materials techniques for evaluation of attributes of sourdough bread with Kombucha SCOBY
- Establishing strength prediction models for low-carbon rubberized cementitious mortar using advanced AI tools
- Investigating the strength performance of 3D printed fiber-reinforced concrete using applicable predictive models
- An eco-friendly synthesis of ZnO nanoparticles with jamun seed extract and their multi-applications
- The application of convolutional neural networks, LF-NMR, and texture for microparticle analysis in assessing the quality of fruit powders: Case study – blackcurrant powders
- Study of feasibility of using copper mining tailings in mortar production
- Shear and flexural performance of reinforced concrete beams with recycled concrete aggregates
- Advancing GGBS geopolymer concrete with nano-alumina: A study on strength and durability in aggressive environments
- Leveraging waste-based additives and machine learning for sustainable mortar development in construction
- Study on the modification effects and mechanisms of organic–inorganic composite anti-aging agents on asphalt across multiple scales
- Morphological and microstructural analysis of sustainable concrete with crumb rubber and SCMs
- Structural, physical, and luminescence properties of sodium–aluminum–zinc borophosphate glass embedded with Nd3+ ions for optical applications
- Eco-friendly waste plastic-based mortar incorporating industrial waste powders: Interpretable models for flexural strength
- Bioactive potential of marine Aspergillus niger AMG31: Metabolite profiling and green synthesis of copper/zinc oxide nanocomposites – An insight into biomedical applications
- Special Issue on Recent Advancement in Low-carbon Cement-based Materials - Part II
- Investigating the effect of locally available volcanic ash on mechanical and microstructure properties of concrete
- Flexural performance evaluation using computational tools for plastic-derived mortar modified with blends of industrial waste powders
- Foamed geopolymers as low carbon materials for fire-resistant and lightweight applications in construction: A review
- Autogenous shrinkage of cementitious composites incorporating red mud
- Special Issue on AI-Driven Advances for Nano-Enhanced Sustainable Construction Materials
- Advanced explainable models for strength evaluation of self-compacting concrete modified with supplementary glass and marble powders
- Special Issue on Advanced Materials for Energy Storage and Conversion
- Innovative optimization of seashell ash-based lightweight foamed concrete: Enhancing physicomechanical properties through ANN-GA hybrid approach
Artikel in diesem Heft
- Review Articles
- Utilization of steel slag in concrete: A review on durability and microstructure analysis
- Technical development of modified emulsion asphalt: A review on the preparation, performance, and applications
- Recent developments in ultrasonic welding of similar and dissimilar joints of carbon fiber reinforcement thermoplastics with and without interlayer: A state-of-the-art review
- Unveiling the crucial factors and coating mitigation of solid particle erosion in steam turbine blade failures: A review
- From magnesium oxide, magnesium oxide concrete to magnesium oxide concrete dams
- Properties and potential applications of polymer composites containing secondary fillers
- A scientometric review on the utilization of copper slag as a substitute constituent of ordinary Portland cement concrete
- Advancement of additive manufacturing technology in the development of personalized in vivo and in vitro prosthetic implants
- Recent advance of MOFs in Fenton-like reaction
- A review of defect formation, detection, and effect on mechanical properties of three-dimensional braided composites
- Non-conventional approaches to producing biochars for environmental and energy applications
- Review of the development and application of aluminum alloys in the nuclear industry
- Advances in the development and characterization of combustible cartridge cases and propellants: Preparation, performance, and future prospects
- Recent trends in rubberized and non-rubberized ultra-high performance geopolymer concrete for sustainable construction: A review
- Cement-based materials for radiative cooling: Potential, material and structural design, and future prospects
- A comprehensive review: The impact of recycling polypropylene fiber on lightweight concrete performance
- A comprehensive review of preheating temperature effects on reclaimed asphalt pavement in the hot center plant recycling
- Research Articles
- Investigation of the corrosion performance of HVOF-sprayed WC-CoCr coatings applied on offshore hydraulic equipment
- A systematic review of metakaolin-based alkali-activated and geopolymer concrete: A step toward green concrete
- Evaluation of color matching of three single-shade composites employing simulated 3D printed cavities with different thicknesses using CIELAB and CIEDE2000 color difference formulae
- Novel approaches in prediction of tensile strain capacity of engineered cementitious composites using interpretable approaches
- Effect of TiB2 particles on the compressive, hardness, and water absorption responses of Kulkual fiber-reinforced epoxy composites
- Analyzing the compressive strength, eco-strength, and cost–strength ratio of agro-waste-derived concrete using advanced machine learning methods
- Tensile behavior evaluation of two-stage concrete using an innovative model optimization approach
- Tailoring the mechanical and degradation properties of 3DP PLA/PCL scaffolds for biomedical applications
- Optimizing compressive strength prediction in glass powder-modified concrete: A comprehensive study on silicon dioxide and calcium oxide influence across varied sample dimensions and strength ranges
- Experimental study on solid particle erosion of protective aircraft coatings at different impact angles
- Compatibility between polyurea resin modifier and asphalt binder based on segregation and rheological parameters
- Fe-containing nominal wollastonite (CaSiO3)–Na2O glass-ceramic: Characterization and biocompatibility
- Relevance of pore network connectivity in tannin-derived carbons for rapid detection of BTEX traces in indoor air
- A life cycle and environmental impact analysis of sustainable concrete incorporating date palm ash and eggshell powder as supplementary cementitious materials
- Eco-friendly utilisation of agricultural waste: Assessing mixture performance and physical properties of asphalt modified with peanut husk ash using response surface methodology
- Benefits and limitations of N2 addition with Ar as shielding gas on microstructure change and their effect on hardness and corrosion resistance of duplex stainless steel weldments
- Effect of selective laser sintering processing parameters on the mechanical properties of peanut shell powder/polyether sulfone composite
- Impact and mechanism of improving the UV aging resistance performance of modified asphalt binder
- AI-based prediction for the strength, cost, and sustainability of eggshell and date palm ash-blended concrete
- Investigating the sulfonated ZnO–PVA membrane for improved MFC performance
- Strontium coupling with sulphur in mouse bone apatites
- Transforming waste into value: Advancing sustainable construction materials with treated plastic waste and foundry sand in lightweight foamed concrete for a greener future
- Evaluating the use of recycled sawdust in porous foam mortar for improved performance
- Improvement and predictive modeling of the mechanical performance of waste fire clay blended concrete
- Polyvinyl alcohol/alginate/gelatin hydrogel-based CaSiO3 designed for accelerating wound healing
- Research on assembly stress and deformation of thin-walled composite material power cabin fairings
- Effect of volcanic pumice powder on the properties of fiber-reinforced cement mortars in aggressive environments
- Analyzing the compressive performance of lightweight foamcrete and parameter interdependencies using machine intelligence strategies
- Selected materials techniques for evaluation of attributes of sourdough bread with Kombucha SCOBY
- Establishing strength prediction models for low-carbon rubberized cementitious mortar using advanced AI tools
- Investigating the strength performance of 3D printed fiber-reinforced concrete using applicable predictive models
- An eco-friendly synthesis of ZnO nanoparticles with jamun seed extract and their multi-applications
- The application of convolutional neural networks, LF-NMR, and texture for microparticle analysis in assessing the quality of fruit powders: Case study – blackcurrant powders
- Study of feasibility of using copper mining tailings in mortar production
- Shear and flexural performance of reinforced concrete beams with recycled concrete aggregates
- Advancing GGBS geopolymer concrete with nano-alumina: A study on strength and durability in aggressive environments
- Leveraging waste-based additives and machine learning for sustainable mortar development in construction
- Study on the modification effects and mechanisms of organic–inorganic composite anti-aging agents on asphalt across multiple scales
- Morphological and microstructural analysis of sustainable concrete with crumb rubber and SCMs
- Structural, physical, and luminescence properties of sodium–aluminum–zinc borophosphate glass embedded with Nd3+ ions for optical applications
- Eco-friendly waste plastic-based mortar incorporating industrial waste powders: Interpretable models for flexural strength
- Bioactive potential of marine Aspergillus niger AMG31: Metabolite profiling and green synthesis of copper/zinc oxide nanocomposites – An insight into biomedical applications
- Special Issue on Recent Advancement in Low-carbon Cement-based Materials - Part II
- Investigating the effect of locally available volcanic ash on mechanical and microstructure properties of concrete
- Flexural performance evaluation using computational tools for plastic-derived mortar modified with blends of industrial waste powders
- Foamed geopolymers as low carbon materials for fire-resistant and lightweight applications in construction: A review
- Autogenous shrinkage of cementitious composites incorporating red mud
- Special Issue on AI-Driven Advances for Nano-Enhanced Sustainable Construction Materials
- Advanced explainable models for strength evaluation of self-compacting concrete modified with supplementary glass and marble powders
- Special Issue on Advanced Materials for Energy Storage and Conversion
- Innovative optimization of seashell ash-based lightweight foamed concrete: Enhancing physicomechanical properties through ANN-GA hybrid approach