Home Microwave-assisted ultrafine silver nanoparticle synthesis using Mitragyna speciosa for antimalarial applications
Article Open Access

Microwave-assisted ultrafine silver nanoparticle synthesis using Mitragyna speciosa for antimalarial applications

  • Vanessa Darakai , Chuchard Punsawad , Jitrayut Jitonnom , Mudtorlep Nisoa and Parawee Rattanakit EMAIL logo
Published/Copyright: April 4, 2024
Become an author with De Gruyter Brill

Abstract

This study investigates the antiplasmodial activity of ultrafine silver nanoparticles (AgNPs, 2–5 nm) synthesized using a green approach involving the Mitragyna speciosa extract and emphasizing the microwave-assisted irradiation technique. Various synthesis parameters were optimized, resulting in the successful production of spherical AgNPs, which exhibited a characteristic surface plasmon resonance peak at around 440 nm. The synthesized AgNPs demonstrated high stability, indicated by a zeta potential value of −28 mV. The antimalarial efficacy of the microwave-assisted AgNPs against the P. falciparum strain was evaluated, demonstrating a half-maximum inhibitory concentration (IC50) value of 1.56 µg·mL−1. Further enhancement in the antimalarial performance was observed when the AgNPs were conjugated with chloroquine (CQ), a traditional antimalarial drug, achieving an impressive IC50 value of 24 ng·mL−1. Additionally, all formulations exhibited low toxicity, with a cytotoxic concentration (CC50) exceeding 800 µg·mL−1 in Vero cells. Complementing these experimental findings, specific computational studies offered insights into the interactions between silver atoms and bioactive compounds in M. speciosa, as well as shedding light on the dynamics of CQ functionalization. These experimental and computational findings emphasize the potential of a sustainable, low-toxicity, and cost-effective AgNP synthesis process, showcasing significant promise in advancing green nanotechnology for the development of effective antimalarial medications.

Graphical abstract

1 Introduction

Malaria, a significant global health challenge caused by various Plasmodium species, continues to impact high-fatality regions such as Southeast Asia, India, the Pacific Islands, and Africa [1,2]. In 2021, the World Health Organization reported approximately 619,000 malaria-related fatalities [3], underscoring the urgent need for more effective and accessible antimalarial therapies. Chloroquine (CQ) or 4-aminoquinoline derivative, a long-standing treatment, faces challenges due to the emergence of CQ resistance [4], prompting the search for innovative therapeutic alternatives.

Silver nanoparticles (AgNPs), known for their synthesis through various routes, including physical, chemical, and notably biological methods, offer a promising solution in numerous applications. Biological methods, especially those using plants, stand out for their cost-effectiveness, ease of implementation, rapidity, non-toxicity, reproducibility, and environmentally friendly nature. These methods effectively circumvent the limitations of traditional nanoparticle synthesis, such as the use of toxic agents and high-energy consumption found in chemical and physical methods. By eliminating these requirements, the biological synthesis of AgNPs offers a more sustainable and eco-friendly approach [5,6]. Various plant extracts, extensively reviewed in the literature [7,8], have been effective in synthesizing AgNPs of various shapes and sizes. Bioactive constituents in these extracts, such as polyphenols, flavonoids, tannins, and terpenoids, play a crucial role in the green synthesis of AgNPs, transforming silver ions (Ag+) into silver atoms (Ag0) through the reduction properties of the molecules found in the phytochemicals. Consequently, the sustainable synthesis of AgNPs using plant extracts is particularly noteworthy for its environmental friendliness and cost-effectiveness.

Temperature is a critical factor in the synthesis of AgNPs, significantly influencing their size and uniformity. Higher temperatures typically result in smaller, more uniform AgNPs, making precise temperature control crucial [5,9]. In this context, microwave irradiation emerges as a favorable technique due to its energy efficiency and capability for uniform and rapid heating, aligning with the principles of green chemistry. This non-combustion method not only reduces carbon emissions compared to conventional heating methods, which often face challenges such as uneven heating and longer processing times but also aims to lower energy consumption and minimize waste. Furthermore, the significant advantage of microwave irradiation lies in its ability to rapidly transfer energy through radiation, as opposed to conventional heat transfer or convection. This ensures fast and instantaneous penetration of energy into materials transparent to microwave radiation, thereby offering the potential for large-scale production of AgNPs [6,10,11,12], a process that is facilitated by easier control and rapid synthesis. These advantages of using microwave irradiation for synthesizing various metallic nanoparticles, including AgNPs, have been well-documented in the review literature [13].

In the current study, we utilize Mitragyna speciosa (M. speciosa), a traditional medicinal plant known for its analgesic properties and compounds, such as mitragynine and 7-hydroxymitragynine [14,15,16], as green reducing and stabilizing agents in the formation of AgNPs. We explore and compare the efficacy of both conventional and microwave-assisted heating techniques in the AgNP synthesis process. Significantly, the AgNPs synthesized using microwave irradiation are further conjugated with CQ, marking a pivotal advancement in developing alternative malaria therapies. Complementing our experimental findings, computational modeling provides deeper insights into the stability and interactions of functional groups within the Ag compounds, supporting the potential of this approach in malaria treatment research.

2 Methodology

2.1 Chemicals and instrumentations

The CQ-resistant P. falciparum (K1) strain was procured and revived at the Research Institute for Health Sciences, Walailak University, Thailand. Gentamicin and nitroblue tetrazolium/phenazine ethosulfate (NBT/PES) were obtained from Sigma-Aldrich, India. 4-(2-Hydroxyethyl)-1-pipera-zineethane sulfonic acid (HEPES) was sourced from Himedia, India. Dimethyl sulfoxide (DMSO) was purchased from Merck, Germany, while AlbuMAX II Lipid-Rich BSA was obtained from Thermo Fisher Scientific, USA. Vero cells were acquired from Elabscience, China. Dulbecco’s modified Eagle’s medium and 2.5% trypsin-EDTA were obtained from Gibco, USA. Fetal bovine serum (10%) and phosphate-buffered saline (PBS) were from Sigma-Aldrich, India. The Roswell Park Memorial Institute-1640 (RPMI-1640) solution was obtained from Gibco, Carlsbad, CA, USA. M. speciosa tea powder was locally sourced from a supermarket in Nakhon Si Thammarat, Thailand, in 2022.

For the experimental setup, an in-house microwave system was developed and facilitated by the Plasma and Electromagnetic Wave Science Center of Excellence (PEwave Center) at the Division of Physics, School of Science, Walailak University, Thailand. An Olympus phase-contrast inverted microscope (Model CK X31) from Hicksville, NY, USA, was utilized to observe cell growth and morphology. The ultraviolet-visible (UV-Vis) spectrophotometer from JASCO (Model V-630), equipped with a standard quartz cuvette from Hellma Analytics (10 mm × 10 mm dimensions), was used to measure the AgNP spectra, scanning between 300 and 600 nm. Insights into the morphology of AgNPs were acquired using the transmission electron microscope (TEM) from JEOL (Model JEM-2100Plus, USA). The stability of the nanoparticles was assessed with a zeta potential analyzer from Brookhaven (ZetaPlus model). An energy-dispersive X-ray (EDX) spectrometer (Oxford’s Aztec model) was used to confirm the elemental composition. The crystalline nature of the AgNPs was determined using the X-ray diffraction (XRD) patterns using a Rigaku SuperNova X-ray diffractometer, set at 50 kV and 0.8 mA, spanning a 2θ angle from 30° to 80°. The Perkin Elmer (Avio 200) inductively coupled plasma-optical emission spectrometry (ICP-OES) instrument was used to measure elemental concentrations. Finally, the Fourier transform infrared (FTIR) spectroscopy tool, the Bruker Tensor27 FTIR spectrometry based in the USA, was used to identify the phytochemical components and functional groups of AgNPs.

2.2 Microwave-assisted green synthesis of AgNPs using M. speciosa extract

To prepare an aqueous extract of M. speciosa, 1.0 g of M. speciosa tea powder was mixed with 50 mL of distilled water, creating a 2% w/v concentration. This mixture was stirred consistently at 60°C for 30 min. The pale-yellow extract was then filtered using Whatman No. 1 filter paper and stored at 4°C for future use.

For the synthesis of AgNPs using microwave irradiation, referred to as AgNPs-MW in this study, 40 mL of a 5 mM AgNO3 solution was added dropwise to 10 mL of the 2% w/v M. speciosa extract. The pH level of the mixture was adjusted to 8 using 0.1 M NaOH and 0.1 M HCl. The solution was then subjected to microwave irradiation for 90 s at a power level of 600 W. A visible shift in color from colorless (indicative of the AgNO3 precursor) to deep brown confirmed the successful reduction of Ag+ ions to Ag0, signaling the formation of AgNPs. After synthesis, the AgNPs-MW colloidal suspension was centrifuged at 25,000 rpm for 60 min to collect the AgNPs-MW. The separated AgNPs-MW were rinsed three times with deionized water to remove residual impurities. To confirm the complete removal of AgNO3 and plant extracts, the supernatants were analyzed using UV-Vis spectrophotometry at 300 nm for AgNO3 and at 203 nm for the plant extracts, which verified the absence of unreacted AgNO3 and residual plant extracts. Finally, the AgNPs-MW were dried at 65°C for 30 min, cooled to ambient temperature, and stored in a vial under appropriate conditions for subsequent use.

2.3 CQ functionalization

To functionalize the AgNPs-MW with CQ, 30 mg of AgNPs-MW were first dispersed in 10 mL of DMSO. Concurrently, 30 mg of CQ was dissolved in 10 mL of PBS solution. The AgNPs-MW colloidal solution was then gradually added dropwise to the CQ solution while maintaining continuous stirring at room temperature for 1 h. Following the stirring process, the mixed solution was centrifuged at 5,500 rpm for 15 min to pellet the conjugated nanoparticles. The resulting AgNPs were then washed three times with DMSO to remove any unbound CQ, ensuring the purity of the nanoparticles. The absence of absorption peaks characteristic of free CQ in supernatants confirmed the successful removal of unbound CQ molecules. After this purification process, the CQ-functionalized AgNPs, now termed AgNPs-MW-CQ, were dried at 65°C for 30 min. Once cooled to ambient temperature, the AgNPs-MW-CQ were stored in a vial for further application.

2.4 Characterization

In the characterization phase, the instruments detailed in Section 2.1 were used to thoroughly analyze the synthesized AgNPs. The UV-Vis spectrophotometer was used to capture their spectral properties, while the TEM revealed insights into their morphology and structural features. The zeta potential analyzer provided stability data, and the EDX spectrometer was used to confirm their elemental composition. Crystalline characteristics were determined via XRD, and the ICP-OES was used to measure elemental concentrations. Finally, FTIR spectroscopy was used to identify the functional groups in the phytochemical compounds and the AgNPs, highlighting their dual roles as reducers and stabilizers.

2.5 Applications

2.5.1 Antimalarial activity

P. falciparum K1, a CQ-resistant strain, was cultured using a synchronous method based on established methods [17], albeit with modifications. The parasite-infested blood was thawed using a solution mixture of 12% NaCl (200 µL) and 1.6% NaCl (10 mL), subsequently washed three times with RPMI media. The parasites were cultivated in a T-25 flask containing 2% human AB-blood cells in a foundational medium of RPMI-1640, supplemented with 0.5% Albumax II. This setup was maintained in a 5% CO2 atmosphere at 37°C. Parasite growth stages and concentrations were monitored using Giemsa-stained thin blood slides and observed under a light microscope at 100× magnification. Upon achieving 10% parasitemia, the culture was transferred to a T-75 flask, ensuring a sustained ring phase under optimal conditions. Evaluation of the antiplasmodial efficacy of all AgNPs was conducted in vitro using a modified parasite lactate dehydrogenase (pLDH) technique, as mentioned in the study of Phuwajaroanpong et al. [17]. This assessment involved a 96-well tissue culture plate. Samples of the raw extract, AgNPs synthesized using the conventional heating method (AgNPs-H), AgNPs-MW, and AgNPs-MW-CQ were dissolved in DMSO. Serial two-fold dilutions from an initial concentration of 100 µg·mL−1 were prepared, with AgNPs-MW-CQ undergoing 12 dilutions. Infected red blood cells (iRBC) were introduced into each well, achieving 2% hematocrit and 2% parasitemia levels. Each well was treated with 1 µg·mL−1 of the corresponding test compound. CQ, with concentrations ranging between 3 and 5 ng·mL−1, acted as the positive control, while DMSO served as the negative control. This procedure was replicated three times. The plates were incubated in a 5% CO2 environment at 37°C for 72 h, subsequently frozen at −80°C for 1 h, and thawed using a water bath for 30 min, resulting in the lysis of the red blood cells. In a new 96-well plate, Malstat reagent (100 µL) was mixed with 20 µg·mL−1 iRBC. Each well was supplemented with 20 µg·mL−1 NBT/PES and incubated for 1 h. Upon completing the pLDH test, absorbance was measured at 620 nm using a microplate reader. In the final step, the percentage of growth inhibition and the half-maximal inhibitory concentration (IC50) were calculated using the following formula:

(1) % Inhibition = 100 × OD of negative control OD of sample OD of negative control

where the OD of the negative control represents DMSO/PBS absorbance, while the OD of the sample pertains to the absorbance of test preparations such as raw extract, AgNPs-H, AgNPs-MW, and AgNPs-MW-CQ.

2.5.2 Cytotoxicity test

For cytotoxicity assessment, the test samples – including M. speciosa raw extract, AgNPs-H, AgNPs-MW, and AgNPs-MW-CQ – were prepared in concentrations ranging from 1.56 to 800 µg·mL−1 through a successive two-fold dilution method. Vero cell cultures were seeded at a density of 1.5 × 104 cells per well, utilizing a 200 µL culture medium in a 96-well plate. The cells were then incubated at 37°C for 24 h in a 5% CO2 atmosphere to reach confluence. Various concentrations of the prepared samples (1 µL each) were introduced to the Vero cell with doxorubicin at final concentrations between 1 and 10 µg·mL−1, serving as the positive control. DMSO and PBS functioned as negative controls. After the addition of samples, cells were incubated for another 48 h at 37°C. This procedure was replicated three times to ensure consistent results. Post-incubation, all wells received MTT reagent and were further incubated for 2 h at 37°C. The MTT reagent was subsequently removed, and each well was treated with 100 µL of DMSO. Absorbance levels were measured at 560 and 670 nm using a microplate reader. The cytotoxicity percentage was calculated with Eq. 2.

(2) % Cytotoxicity = 100 100 × OD of the sample well OD of the negative control well

where the OD of the negative control signifies the DMSO/PBS absorbance, while the OD of the sample pertains to the absorbance of preparations like raw extract, AgNPs-H, AgNPs-MW, and AgNPs-MW-CQ.

2.6 Computational study

To explore the interaction between a silver atom and M. speciosa-derived compounds, quantum chemistry calculations were carried out using the density functional theory (DFT) method [18,19,20]. Each of the five secondary metabolites (SMs) – alkaloid, flavonoid, terpenoid, tannin, and saponin – and CQ were examined for silver interactions. All of the structures, including the silver atom, metabolites, and complexes (AgNPs) formed between them, were optimized in the gas phase using Gaussian 09 software [21]. The calculations were carried out using Becke’s three-parameter functional combined with the Lee–Yang–Parr hybrid functional, known as the B3LYP level [18,19]. The B3LYP functional has been previously used to study AgNPs [22,23] with success in terms of stable geometry and reliable results [24,25]. For the silver atom, the LANL2DZ basis set [26] was utilized, while the bioactive molecules were treated with the 6–31+G(d, p) basis set. This B3LYP/mixed basis set approach has previously been applied successfully in the literature [27,28,29].

The Gibbs interaction energy, ΔG int, between a silver atom and each SM molecule was calculated as follows:

(3) G int = G Ag_SM ( G Ag + G SM )

where G Ag_SM , G Ag , and G SM are the Gibbs free energies of the corresponding Ag_SM complexes, one silver atom as Ag(0) neutral species, and the SM, respectively. Negative values of G int correspond to favored metal-SM binding. A similar procedure was also applied to the CQ drug.

A visual summary of the proposed work can be found in Figure 1.

Figure 1 
                  Schematic representation of the concept of this work.
Figure 1

Schematic representation of the concept of this work.

3 Results and discussion

3.1 Preliminary results

In this section, we present two sets of preliminary results: phytochemical screening of M. speciosa extract and optimization of synthesis parameters for AgNPs-MW. These initial investigations were conducted to establish a foundational understanding of the key components in the M. speciosa extract responsible for the reduction and stabilization of AgNPs, as well as to determine the optimal conditions for the microwave-assisted synthesis process.

3.1.1 Phytochemical screening of M. speciosa extract

The phytochemical constituents present in the M. speciosa extract were determined using standard methods, as described in previous literature [30]. This qualitative screening was essential to verify the presence of specific phytochemicals in the extract, notably flavonoids, tannins, alkaloids, terpenoids, and saponins. The presence of these compounds in M. speciosa aligns with findings reported in the existing literature [14,15,16], where Mitragynine is highlighted as the main alkaloid in the extract, along with other compounds such as flavonoids and phenolic acids. These constituents were identified using several techniques, including GC-MS, LC-MS/MS, and UPLC-PDA, demonstrating the comprehensive phytochemical profile of the extract. Additionally, in this work, the total phenolic contents of M. speciosa’s aqueous extract were determined quantitatively using the Folin-Ciocalteu reagent. With gallic acid as the standard, measured at 760 nm, the contents were found to be 146.7 mg GAE/g, providing further insight into the extract’s phytochemical profile. For a detailed account of the specific phytochemical identified, refer to Table S1 in the Supporting Information (SI) section. It is crucial to emphasize the role of these identified phytochemicals in the synthesis of AgNPs. They are instrumental in the reduction of Ag+ ions to Ag0 atoms and play a significant role in stabilizing the synthesized nanoparticles.

3.1.2 Optimization of synthesis parameters for AgNPs-MW

The synthesis of AgNPs necessitates meticulous consideration of various parameters, as these significantly influence the nanoparticles’ characteristics, including size, morphology, and yield. Typically, the successful synthesis of AgNPs is confirmed by observing the surface plasmon resonance (SPR) band through UV-Vis spectroscopy. A distinct SPR band within approximately 350–450 nm is indicative of AgNP formation [30,31]. The position, sharpness, and width of the SPR spectra are crucial, as they provide insights into the size distribution and morphology of the AgNPs [32,33].

In this study, we undertook a comprehensive optimization of synthesis conditions, exploring both conventional heating and microwave irradiation techniques. The research utilized M. speciosa extract, which played a dual role in the AgNP synthesis, serving as both a reducing and stabilizing agent. We carefully adjusted various parameters, including silver nitrate concentration, extract concentration, pH, reaction time, and temperature, to optimize the synthesis conditions. Our exploration began with establishing a foundational synthesis protocol using conventional heating on a hotplate. This protocol was then adapted for the microwave irradiation approach, employing an in-house microwave system. The primary focus of the study was to compare the efficacy of these two heating methods in synthesizing AgNPs, with an emphasis on the advantages offered by microwave-assisted synthesis in terms of process efficiency and nanoparticle quality.

3.1.2.1 Conventional heating method (AgNPs-H)

In the conventional heating method, the pH level of the reaction system was identified as a critical factor influencing the formation of AgNPs. Adjusting the pH affects the charge on metabolites, subsequently influencing the redox reduction process and the interaction between silver ions and phytochemical mediators. A comprehensive investigation across a pH range of 2 to 10, detailed in Figure S1(a), revealed that acidic conditions hindered AgNP formation, whereas basic conditions facilitated it. The optimal pH level was identified at 8.

The concentration of the silver ion precursor, AgNO3, emerged as another pivotal parameter. Explorations across concentrations from 1 to 5 mM showed a clear correlation: higher AgNO3 concentration resulted in an increased intensity of the SPR peak. The optimal concentration was established at 5 mM, as higher concentrations led to undesirable effects like the silver mirror phenomenon (Figure S1(b)).

Additionally, the concentration of M. speciosa extract, serving as a reducing and stabilizing agent, varied from 2% to 5% (w/v). Higher extract concentrations were found to inhibit AgNP formation [34], with 2% (w/v) identified as optimal, balancing inhibitory effects and particle size (Figure S1(c)).

The volume ratio of AgNO3 precursor solution to M. speciosa extract also proved crucial, directly influencing the yield and size of the nanoparticles. A ratio of A 4:1 demonstrated superior results in terms of yield and size distributions, as indicated by higher intensity and sharper spectral bands in the UV-Vis spectra (Figure S1(d)).

Finally, the synthesis reaction time and temperature using the conventional heating method were investigated. A reaction time of 90 min at 90°C was established as optimal, providing an effective balance between reaction efficiency and sensitivity. This condition resulted in a noticeable enhancement in SPR intensity, as shown in Figure S1(e and f) of the SI.

3.1.2.2 Microwave irradiation (AgNPs-MW)

In the microwave-assisted synthesis of AgNPs, two additional factors, reaction period and microwave irradiation power, were investigated and benchmarked against the optimal conditions identified through the conventional heating method. The optimization involved varying the reaction time from 0 to 150 s at 30 s intervals, maintaining a constant microwave power of 600 W. Figure S2(a) illustrates the impact of the reaction time on the SPR bands. It was observed that the intensity of the reaction mixture progressively darkened with increasing reaction time up to 90 s, indicative of a rise in AgNPs yield. However, beyond 90 s, a slight drop in intensity occurred, establishing 90 s as the optimal reaction duration for the highest AgNP yield. Regarding microwave irradiation power, which varied from 300 to 1,000 W, the SPR absorbance increased with power up to 600 W and decreased significantly at higher levels. The maximum peak absorbance was observed at 600 W, as shown in Figure S2(b). Thus establishing it as the optimum microwave irradiation power. To confirm the role of plant extract in synthesizing AgNPs under optimal conditions, additional experiments were conducted using AgNO3 in various types of media (DI water, NaOH solution, plant extract solution, and a mixture of plant extract in NaOH solution). The results, as depicted in Figure S3, indicate that the plant extract, particularly under alkaline conditions, is crucial for the reduction of Ag+ ions and the formation of AgNPs through microwave irradiation.

The ranges and optimal values for each heating method are summarized in Table 1. Compared to conventional methods, microwave irradiation stands out, reducing the reaction time dramatically from 90 min (with conventional heating at 90°C) to just 90 s (at 600 W) while maintaining optimal conditions for nanoparticle synthesis. This rapid reaction rate, facilitated by microwave irradiation, promotes homogeneous nucleation and efficient nanoparticle synthesis [35]. Thus, integrating microwave chemistry with biosynthetic methods significantly enhances nanoparticle synthesis, adhering to green chemistry principles. The remarkable time efficiency of microwave-assisted synthesis offers significant advantages for large-scale production, underscoring its effectiveness and potential in nanotechnology applications.

Table 1

Comparison of the optimal conditions for synthesis of AgNPs using conventional heating and microwave irradiation methods

Method Studied range Optimal conditions
Conventional heating
pH 2–10 8
Concentration of AgNO3 (mM) 1–5 5
Extract concentration (%w/v) 1–5 2
Volume ratio (silver salt to extract) 1:4–4:1 4:1
Heating time (min) 30–120 90
Temperature (°C) 40–90 90
Microwave heating
Time (min) 0–2 1.5
Power (W) 300–1,000 600

3.2 Characterization of the synthesized AgNPs using M. speciosa extract

During the synthesis of AgNPs using M. speciosa extract for both reduction and stabilization, a notable color transition was observed. The solution changed from its original colorless state, characteristic of the AgNO3 solution, or a light-yellow hue, typical of the extract solution, to a distinct dark brown. This color change is a well-recognized indicator of the successful formation of AgNPs. The SPR bands were distinctly observed at different wavelengths for the AgNPs synthesized using the two methods. For AgNPs synthesized by conventional heating (AgNPs-H), the SPR band was observed at 440 nm, as shown in Figure 2a. In contrast, AgNPs synthesized by microwave irradiation (AgNPs-MW) exhibited an SPR band at 415 nm, as depicted in Figure 2b. The observed blue shift in the SPR band for AgNPs-MW, compared to AgNPs-H, suggests that the nanoparticles produced by the microwave irradiation method are smaller. This size difference, indicated by the shift of the SPR band aligns with existing literature on nanoparticle characterization [36,37].

Figure 2 
                  UV-Vis absorption spectra accompanied by digital photographs of the AgNPs synthesized using M. speciosa extract under two distinct heating techniques: (a) conventional heating and (b) microwave irradiation.
Figure 2

UV-Vis absorption spectra accompanied by digital photographs of the AgNPs synthesized using M. speciosa extract under two distinct heating techniques: (a) conventional heating and (b) microwave irradiation.

TEM micrographs, as presented in Figure 3, reveal the spherical morphology of the AgNPs synthesized using both heating methods. For AgNPs-H, a size range of 8 to 17 nm was observed, as depicted in Figure 3a. In contrast, AgNPs-MW displayed a significantly reduced size range of 2–4 nm, as depicted in Figure 3b. This observation is in line with the earlier noted blue shift in the SPR band, further confirming the size reduction of AgNPs when synthesized using microwave irradiation. The more uniform and smaller size distribution achieved under microwave irradiation can be attributed to its capability for rapid and homogeneous heating. This heating method accelerates the reaction kinetics, facilitating quicker nucleation and growth of AgNPs. Consequently, this expedited process leads to the formation of smaller nanoparticles compared to those obtained through slower conventional heating methods. The difference in the size of AgNPs, as observed in TEM images, corroborates with the literature on the effects of synthesis methods on nanoparticle characteristics [38,39,40].

Figure 3 
                  TEM micrographs and size distribution showing the morphology of the AgNPs-H (a) and AgNPs-MW (b) using M. speciosa extract.
Figure 3

TEM micrographs and size distribution showing the morphology of the AgNPs-H (a) and AgNPs-MW (b) using M. speciosa extract.

The EDX spectra for both AgNPs-H and AgNPs-MW confirm the presence of silver, characterized by peaks at approximately 3 keV, as evidenced in Figures S4(a–b). This peak is a distinctive indicator of silver in the nanoparticles. Zeta potential analyses reveal values of −21.12 mV for AgNPs-H and −28.63 mV for AgNPs-MW. These results indicate that AgNPs-MW exhibits greater colloidal stability compared to those AgNPs-H. This enhanced stability is likely attributable to the rapid and uniform heating provided by microwave irradiation at the optimum power level. Such irradiation influences the rate of nucleation and growth of nanoparticles, thereby affecting their size distribution and stability. Unlike conventional methods, microwave irradiation ensures even energy distribution throughout the reaction mixture, preventing localized overheating and the formation of hot spots, which are common causes of particle agglomeration. Consequently, the nanoparticles remain well-dispersed, contributing to their increased stability, as supported by the literature [11].

The XRD spectra, illustrated in Figure 4, confirm the crystalline nature of the synthesized nanoparticles. AgNPs-H exhibit diffraction peaks at 2θ angles of 38.01, 44.37, 64.46, and 77.36°, corresponding to the (111), (200), (220), and (311) planes of face-centered cubic (fcc) silver. Similarly, AgNPs-MW shows diffraction peaks at 2θ angles of 38.36, 44.38, 64.76, and 77.53°. These peaks also align with the (111), (200), (220), and (311) planes of fcc silver, confirming the successful synthesis and crystalline structure of the AgNPs. These findings are in agreement with established literature [41].

Figure 4 
                  XRD pattern of the AgNPs obtained by both synthesized heating methods: AgNPs-H (a) and AgNPs-MW (b).
Figure 4

XRD pattern of the AgNPs obtained by both synthesized heating methods: AgNPs-H (a) and AgNPs-MW (b).

The silver content of the synthesized AgNPs was quantified using ICP-OES. The analysis revealed a notable difference in the silver composition of the nanoparticles produced by the two methods. AgNPs-H comprised 68% silver by weight, while AgNPs-MW contained a significantly higher silver percentage of 88% (w/w). This disparity underscores the superior efficacy of microwave irradiation in enhancing reaction kinetics and heat transfer, resulting in a higher yield of AgNPs. The advantages of microwave irradiation as a synthesis technique are thus clearly demonstrated. It offers increased energy efficiency, improved product yield, accelerated reaction rates, and high-quality nanoparticle production with minimal agglomeration. The deployment of microwave irradiation in nanoparticle synthesis aligns well with the principles of green chemistry, emphasizing its potential for environmentally sustainable applications. This technique’s ability to efficiently produce nanoparticles with a higher silver content highlights its utility and superiority in the field of nanoparticle synthesis, particularly for applications requiring high purity and efficiency.

The FTIR spectra of AgNPs-H and AgNPs-MW were compared with those of the M. speciosa extract, as illustrated in Figure 5. In the AgNPs-H spectra, characteristic absorption bands were identified at specific wavenumbers: 3,423 cm−1, which corresponds to the O–H stretching of hydroxyl groups; 2,919 cm−1, representing C–H stretching in aromatic structures; and 1,606 cm−1, indicative of C═O vibrations; and 1,059 cm−1, associated with C–O vibrations. Similarly, in the AgNPs-MW spectra, these bands were observed at slightly shifted wavenumbers: 3,406, 2,912, 1,600, and 1,057 cm−1, respectively. The observed redshift in the vibrational frequencies for both sets of AgNPs, compared to the M. speciosa extract, suggests an interaction between the functional groups of the extract and the AgNPs’ surfaces. The FTIR results not only confirm the successful synthesis of the nanoparticles but also highlight the pivotal role played by the bioactive molecules from the extract in the synthesis process. This shift in wavenumber demonstrates the chemical modifications and interactions at the nanoparticle surfaces, providing crucial insights into the nature of the AgNPs synthesized [42].

Figure 5 
                  FTIR spectra comparison across different samples: M. speciosa crude powder (a), AgNPs synthesized via conventional heating (AgNPs-H) (b), and AgNPs produced using microwave irradiation (AgNPs-MW) (c).
Figure 5

FTIR spectra comparison across different samples: M. speciosa crude powder (a), AgNPs synthesized via conventional heating (AgNPs-H) (b), and AgNPs produced using microwave irradiation (AgNPs-MW) (c).

Polyphenolic compounds present in M. speciosa extracts play a pivotal role in the biosynthesis of AgNPs, serving dual functions as both reducing and stabilizing agents. During the initial phase of biosynthesis, these polyphenolic compounds donate electrons to silver ions (Ag+). This donation triggers a reduction reaction that converts the ions into metallic AgNPs (Ag0). This reduction process is efficiently facilitated by the inherent properties of polyphenols, which act as potent electron donors [43]. Following the reduction phase, these same polyphenolic molecules assume a critical role in stabilization. They adsorb onto the nascent surfaces of the AgNPs, primarily through interactions involving their hydroxyl and carboxyl functional groups. This adsorption forms a protective layer around the nanoparticles, effectively preventing aggregation. It provides steric and electrostatic barriers, thereby significantly enhancing the colloidal stability and dispersity of the nanoparticles [44,45]. Through these simultaneous processes of reduction and stabilization, polyphenolic compounds from M. speciosa extracts enable the synthesis of AgNPs that are not only stable but also well-dispersed. This dual functionality is essential for the effective production of AgNPs and highlights the utility of natural extracts in green nanotechnology applications. The schematic illustrating the plausible mechanism for AgNP formation is provided in Figure S5 of the SI.

To elucidate the nature of the interactions between silver atoms and the phytochemicals present in the extract, including flavonoids, tannins, alkaloids, terpenoids, and saponins, DFT calculations were employed. DFT is a quantum mechanical method used to investigate the electronic structures of molecules and their interactions with metals. It is instrumental in predicting which natural compounds might be most effective in reducing and stabilizing nanoparticles. As illustrated in Figure 6, the DFT study highlighted tannins as having the strongest affinity for silver atoms. This is evidenced by the shorter bond lengths observed between the silver atom and the oxygen atoms of the tannin’s exocyclic carbonyl group, measuring 2.31 and 2.26 Å. These distances indicate stronger interactions, corroborated by the most negative Gibbs free energy of interaction (ΔG int) value of −17.5 kcal·mol−1. Such favorable interaction energies suggest that tannins are likely to play a significant role in the stabilization of AgNPs, possibly through chelation, where multiple bonds form between the tannin and the silver atom. Alkaloids also demonstrated a potential for interaction with silver atoms, as indicated by a bond length of 2.27 Å and a ΔG int of −8.0 kcal·mol−1, pointing to their possible involvement in the synthesis process, albeit to a lesser extent than tannins. In contrast, interactions between silver atoms and compounds such as saponins, terpenoids, and flavonoids were less favorable. This is reflected by longer bond distances and positive ΔG int values. These positive interaction energies suggest endothermic reactions, which are thermodynamically non-spontaneous under standard conditions. Consequently, these compounds are less likely to contribute significantly to the formation of stable AgNPs.

Figure 6 
                  Most stable geometries and interaction energies (ΔG
                     int) of the one-silver atom (shown in cyan color) complexed with tannin (a), alkaloid (b), saponin (c), terpenoid molecules (d), and flavonoid (e). The enthalpic values of the interaction energies, ΔH
                     int, are also included in parenthesis. The difference between ΔG
                     int and ΔH
                     int can reflect the entropic effect for each of the Ag-bioactive compound systems.
Figure 6

Most stable geometries and interaction energies (ΔG int) of the one-silver atom (shown in cyan color) complexed with tannin (a), alkaloid (b), saponin (c), terpenoid molecules (d), and flavonoid (e). The enthalpic values of the interaction energies, ΔH int, are also included in parenthesis. The difference between ΔG int and ΔH int can reflect the entropic effect for each of the Ag-bioactive compound systems.

The DFT calculations extended beyond Gibbs free energy to include an analysis of the enthalpic interaction energies (ΔH int), which further supported and corroborated the trends observed with the trend observed with ΔG int. For tannins and alkaloids, which had more negative ΔG int values, negative ΔH int values were also observed. This indicates that their interactions with silver atoms are not only spontaneous but also exothermic. The release of heat in these interactions points to the formation of a stable complex between these phytochemicals and silver atoms. Conversely, the positive ΔH int values associated with saponins, terpenoids, and flavonoids suggest that interactions involving these compounds are less energetically favorable. This finding aligns with the thermodynamic prediction that these compounds are less effective in stabilizing AgNPs. The positive ΔH int values indicate that the interactions are endothermic, requiring energy input and thus less likely to contribute to the stable formation of AgNPs under standard conditions. The DFT calculations, through both ΔG int and ΔH int analyses, provide a comprehensive understanding of the molecular interactions at play in the biosynthesis of AgNPs. They elucidate the role of specific phytochemicals in the reduction and stabilization processes, highlighting the critical function of tannins and alkaloids in forming stable AgNP complexes.

3.3 In vitro antiplasmodial and cytotoxic activities

3.3.1 AgNPs without CQ conjugation

In assessing the antiplasmodial efficacy of the synthesized AgNPs, the IC50 values are crucial for evaluating their potency. These values, determining the concentration required to inhibit 50% of the parasitic activity, are essential in quantifying the extent to which these nanoparticles can suppress malaria-causing parasites. IC50 values below 5 µg·mL−1 are considered highly potent, indicating strong antiplasmodial agents. Those within the range of 5–15 µg·mL−1 are deemed to have good antiplasmodial activity, suitable for further development. Conversely, values between 15 and 50 µg·mL−1 may only confer moderate activity, necessitating improvements in nanoparticle design or combination with other agents to enhance efficacy. Values exceeding 100 µg·mL−1 suggest limited therapeutic potential against malaria. The determination of these values for both AgNPs-H and AgNPs-MW is a critical step in the preclinical development of nanoparticle-based antimalarial therapies [17].

Table 2 presents the comparative antiplasmodial activities of the crude extracts of M. speciosa, AgNPs-H, and AgNPs-MW against the P. falciparum K1 strain. Notably, the AgNPs-H displayed IC50 values of 6.25 ± 0.99, categorizing them as having good antiplasmodial activity. In contrast, the AgNPs-MW exhibited a significantly lower IC50 value of 1.56 ± 0.51 µg·mL−1, surpassing the threshold for high potency against the parasite and suggesting an exceptional antiplasmodial potential. The enhanced activity of AgNPs-MW can be attributed to their reduced size range of 2–4 nm, which provides a larger relative surface area. This increased surface area may facilitate more effective interactions with biological targets within the parasite, enhancing uptake and/or binding to key biological molecules, thereby increasing antiplasmodial efficacy. In comparison, the M. speciosa crude extract exhibited an IC50 of 25 ± 0.85 µg·mL−1, falling within the moderate activity range. This underlines the significant enhancement in antiplasmodial properties achieved through the nanoparticle synthesis process beyond the inherent capabilities of the plant extract constituents.

Table 2

Comparison of IC50, CC50, and selective index values among the crude extract, AgNPs synthesized using both heating methods and drug-loaded AgNPs-MW-CQ

Sample IC50 (µg·mL−1) CC50 (µg·mL−1) SIa
M. speciosa crude extract 25.0 ± 0.85 >800 32
AgNPs-H 6.25 ± 0.99 >800 128
AgNPs-MW 1.56 ± 0.51 >800 513
AgNPs-MW-CQ 24 ± 2.85 (ng·mL−1) >800 33,333
CQ (positive control) 3.75 ± 0.31 (ng·mL−1)
Doxorubicin (positive control) 1.25

aSelective index (SI) is calculated as SI = CC50/IC50, where CC50 represents the concentration causing 50% cytotoxicity, and IC50 represents the concentration achieving 50% inhibition or the effective concentration.

The cytotoxicity effects of both the crude extract of M. speciosa and the synthesized AgNPs on Vero cells were thoroughly evaluated, with the findings presented in Table 2. The 50% cytotoxicity concentration (CC50) is used as a benchmark to assess the extent to which a substance can reduce cell viability or proliferation by 50%. Notably, all tested samples displayed CC50 values greater than 800 µg·mL−1, indicating minimal cytotoxic impact on the Vero cell line and suggesting a favorable biocompatibility profile. Such high CC50 values are indicative of a low risk of cytotoxicity, which is desirable for therapeutic applications. For comparison, substances with CC50 values below 30 µg·mL−1 after 48–72 h are generally considered cytotoxic [46]. According to Table 2, the CC50 values for the crude extract, AgNPs-H, and AgNPs-MW all exceeded 800 µg·mL−1. This indicates low cytotoxicity for each of these substances. In contrast, doxorubicin, a conventional cytotoxic agent used as a positive control in this study, exhibited potent cytotoxic activity with a CC50 value of 1.25 µg·mL−1. The marked difference in CC50 values between doxorubicin and the tested samples highlights the sensitivity of the cytotoxicity assay employed and validates the low cytotoxicity profile of the AgNPs and the crude M. speciosa extract. Furthermore, the selectivity index values of the tested samples against P. falciparum ranged from approximately 30–30,000, calculated as CC50/IC50, as suggested by [47]. These values indicate a high degree of selectivity and efficacy, considering that a ratio of ≥10 is generally regarded as indicative of a good therapeutic index for a remedy or drug [48]. The substantial selectivity index values underscore the potential of these AgNPs as bioactive materials with significant therapeutic utility against malaria.

The well-documented antioxidant properties of M. speciosa, rich in phytochemicals known to neutralize free radicals and alleviate oxidative stress-induced disorders, provide a solid foundation for developing safe therapeutic agents [49,50]. The AgNPs derived from M. speciosa extracts demonstrate negligible cytotoxicity toward Vero cells, suggesting their non-toxic nature. This indicates the absence of adverse effects on normal cellular structures, a critical aspect of therapeutic applications. Further bolstering their potential, the in vitro antimalarial and cytotoxicity assays highlight a distinct advantage of AgNPs synthesized using microwave irradiation. These AgNPs-MW not only exhibit superior antiplasmodial potency but also maintain a high degree of biocompatibility. Such attributes position microwave-synthesized AgNPs as promising candidates for advanced research into antimalarial interventions.

It is important to note that in assessing the antiplasmodial effects of AgNPs, a control experiment was conducted using the solvent alone to ensure its impact on hemolysis was negligible, confirming that the observed antiplasmodial effects were not due to solvent-induced hemolysis. Further analysis revealed that the concentration range of AgNPs exhibiting antiplasmodial activity, from 24 ng·mL−1 to 6.25 µg·mL−1, does not induce hemolysis in red blood cells. This conclusion is consistent with the findings in the study of Luna-Vázquez-Gómez et al. [51], which demonstrate that AgNPs at concentrations up to 24 µg·mL−1 do not cause hemolysis. Thus, the observed antiplasmodial activity is attributed to the inherent efficacy of AgNPs, independent of any hemolytic effects.

3.3.2 AgNPs-MW with CQ functionalization

CQ has long been a cornerstone in treating and preventing malaria [52]. Despite its effectiveness, challenges such as severe toxicity and the emergence of drug-resistant strains have necessitated novel approaches. One such approach is the conjugation of CQ with AgNPs. In this study, CQ was bound to AgNPs-MW, resulting in CQ-functionalized AgNPs-MW (AgNPs-MW-CQ). The functionalization was substantiated through a range of characterization techniques, including UV-visible spectroscopy, EDX analysis, FTIR spectra, and DFT calculations.

The UV-Vis spectra, as shown in Figure 7a, demonstrate the presence of two distinct peaks: one at 343 nm, characteristic of CQ [53], and another at 415 nm, indicative of AgNPs-MW. In the spectrum of AgNPs-MW-CQ, both peaks are present, confirming the successful conjugation. Additionally, a noticeable shift in the plasmon band is observed upon CQ addition, with peaks at 347 and 428 nm, suggesting changes in the nanoparticles’ environment.

Figure 7 
                     Characterization of CQ-functionalized AgNPs (AgNPs-MW-CQ): UV-Vis spectra comparing CQ, AgNPs-MW, and AgNPs-MW-CQ (a), EDX analysis of AgNPs-MW-CQ (b), EDX-layer area representation for AgNPs-MW-CQ, with green regions indicating the silver element and red regions representing nitrogen (c), FTIR spectra for CQ, AgNPs-MW, and AgNPs-MW-CQ of CQ (d).
Figure 7

Characterization of CQ-functionalized AgNPs (AgNPs-MW-CQ): UV-Vis spectra comparing CQ, AgNPs-MW, and AgNPs-MW-CQ (a), EDX analysis of AgNPs-MW-CQ (b), EDX-layer area representation for AgNPs-MW-CQ, with green regions indicating the silver element and red regions representing nitrogen (c), FTIR spectra for CQ, AgNPs-MW, and AgNPs-MW-CQ of CQ (d).

The EDX spectrum (Figure 7b) reveals a distinct silver signal at 3 keV, alongside high-intensity signals for elements associating with CQ (nitrogen, phosphorus, and chloride) at approximately 0.27, 2.03, and 2.67 keV, respectively. Furthermore, EDX mapping (Figure 7c) showcases areas highlighted in green and red, denoting the presence of silver and nitrogen elements. The co-localization of silver from AgNPs and nitrogen from CQ provides robust evidence for the successful functionalization of AgNPs with CQ.

Figure 7d displays the FTIR spectra for AgNPs-MW, CQ, and AgNPs-MW-CQ. The spectra exhibit similar patterns, with O–H vibrational peaks between 2,425 and 2,440 cm−1, C–H stretching resonances near 2,920–2,939 cm−1, C═O bonding within 1,598–1,616 cm−1, C–O bond vibrations between 1,421 and 1,458 cm−1, and C–O stretching peaks around 1,055–1,091 cm−1. Notably, the spectrum of CQ alone displays a characteristic N–H vibrational peak at 3,232 cm−1 [40], which shifts to 3,252 cm−1 in the spectrum for AgNPs-MW-CQ. This shift signals that AgNPs-MW has been successfully modified with CQ and serves as further evidence for the successful functionalization of the nanoparticles.

DFT modeling of both the protonated and deprotonated structures of AgNPs-MW-CQ, as depicted in Figure 8, offers additional insight into the interaction dynamics between the AgNPs and CQ. The modeling results reveal distinct differences in thermodynamic favorability between the two forms of CQ. The interaction between silver and the deprotonated form of CQ is shown to be thermodynamically more favorable, exhibiting a significant negative interaction energy (ΔG int) of −20.3 kcal·mol−1. This indicates that the deprotonated CQ forms a more stable complex with the AgNPs-MW. In contrast, the interaction with the protonated form of CQ is less favorable, as indicated by a positive ΔG int of +16.6 kcal·mol−1. Such positive interaction energy suggests a lower propensity for complex formation. Additionally, the modeling results highlight a notable difference in the Ag–N bond distance between the two forms. In the deprotonated CQ, the Ag–N bond distance is shorter, by 0.6 Å, compared to that in the protonated form. This shorter bond length in the deprotonated state underscores a stronger interaction and enhanced stability when CQ is complexed with the AgNPs in their deprotonated form. These findings from DFT modeling provide crucial molecular-level details, shedding light on the more effective interaction mechanism of deprotonated CQ with AgNPs. The results have significant implications for understanding the stability and efficacy of CQ functionalized AgNPs in potential therapeutic applications.

Figure 8 
                     Illustration of the most stable geometries and interaction energies (ΔG
                        int) for a single-silver atom (shown in cyan) complexed with the CQ drug. The interaction with the protonated nitrogen atom (a) and the interaction with the deprotonated nitrogen atom (b) at the Ag metal site. The colors of the atoms are as follows: silver (cyan), nitrogen (blue), oxygen (red), hydrogen (white), carbon (black), and chloride (green).
Figure 8

Illustration of the most stable geometries and interaction energies (ΔG int) for a single-silver atom (shown in cyan) complexed with the CQ drug. The interaction with the protonated nitrogen atom (a) and the interaction with the deprotonated nitrogen atom (b) at the Ag metal site. The colors of the atoms are as follows: silver (cyan), nitrogen (blue), oxygen (red), hydrogen (white), carbon (black), and chloride (green).

The collective results from the characterizations confirm the successful formation of AgNPs-MW-CQ, supporting the hypothesis that CQ can effectively coat AgNPs. Subsequent evaluations of antiplasmodial and cytotoxic activities, detailed in Table 2, reveal a marked increase in efficacy against the P. falciparum K1 strain following CQ functionalization. The non-coated CQ AgNPs exhibited an IC50 value of 1.56 ± 0.51 µg·mL−1, whereas AgNPs-MW-CQ showed a significantly reduced IC50 of 24 ± 2.85 ng·mL−1, indicating a substantial enhancement in antimalarial potency. Chaves et al. [54] have noted the effectiveness of drug-nanocarrier combinations in inhibiting P. falciparum transmission, where drug-loaded nanoparticles act as targeted delivery vehicles. Moreover, both formulations of AgNPs exhibited CC50 values well above 800 µg·mL−1, demonstrating their low toxicity toward regular cells. These findings highlight the potential of AgNPs-MW-CQ as a novel antimalarial intervention, offering improved therapeutic effects. The enhanced antimalarial potency and low cytotoxicity of AgNPs-MW-CQ position them as promising candidates in the fight against malaria, potentially revolutionizing treatment strategies by utilizing the targeted delivery capabilities of nanocarriers.

It is crucial to note that AgNPs can induce toxicity to some extent in animal or human systems when administered at concentrations above the lowest observed adverse effect level, potentially leading to various health problems. This underscores the importance of understanding their biodistribution, metabolism, excretion, and overall safety profile. The toxicity of AgNPs and their potential mechanisms have been extensively reviewed in the study of Nie et al. [55], which highlights that toxicity is influenced by intrinsic properties such as shape, size, and surface modification, as well as external factors like dosage. Furthermore, research cited in this review demonstrates that AgNPs may accumulate in organs such as the liver, spleen, kidneys, and lungs, potentially causing oxidative stress, inflammation, and cytotoxic effects. These findings underscore the need for comprehensive toxicological evaluations to thoroughly assess the safety of AgNPs. Despite significant advances, knowledge gaps remain in fully understanding the long-term effects of AgNPs, particularly regarding their clearance mechanisms and potential accumulation in the human body. Regulatory authorities have emphasized the necessity of detailed safety assessments for nanomaterials intended for biomedical applications. While this study has primarily focused on the antiplasmodial activity of AgNPs, future research should aim to address these critical aspects of nanotechnology application, ensuring both efficacy and safety in potential therapeutic roles.

4 Conclusions

Addressing the global challenge of malaria, particularly in the face of emerging drug-resistant strains, necessitates innovative and effective therapeutic strategies. This study makes a significant contribution to this crucial endeavor by leveraging the unique bioactive compounds of M. speciosa in the green synthesis of AgNPs. Utilizing microwave irradiation, we have developed a sustainable and efficient approach for producing ultrafine, spherical, and stable AgNPs. The small size and physicochemical properties of these AgNPs are pivotal in enhancing their antimalarial activity. A notable advancement in this research is the integration of these AgNPs with CQ, leading to a significant enhancement in antimalarial efficacy against P. falciparum. This is demonstrated by the substantial reduction in the IC50 value to 24 ± 2.85 ng·mL−1, compared to the standalone AgNPs. Our findings highlight a preferential interaction between the silver atom and the deprotonated form of CQ, shedding light on the molecular dynamics underlying this enhanced efficacy and suggesting the potential of AgNPs to improve antimalarial therapeutic outcomes. Crucially, the AgNP formulations, including the biocompatible AgNPs-MW-CQ, have demonstrated non-toxicity to Vero cells. This positions them as promising candidates for further research and potential therapeutic applications. We advocate for continued research in this field, particularly focusing on in vivo studies and clinical trials to validate the efficacy and safety of AgNPs in antimalarial interventions. In conclusion, our research underscores the potential of M. speciosa-mediated AgNPs, especially when combined with CQ, to offer a novel and potent solution in the ongoing battle against malaria. While the transition from laboratory research to clinical applications presents numerous challenges, our findings pave the way for future innovations and emphasize the critical importance of exploring nature-derived nanoparticles in addressing infectious diseases.

Acknowledgement

The authors extend their heartfelt gratitude to the Functional Materials and Nanotechnology Centre of Excellence at Walailak University (WU) for their generous financial support and the provision of research facilities. Our appreciation also goes to the Research Institute for Health Sciences, WU, Thailand, for their collaboration and for providing the necessary malaria and cell culture facilities. Special thanks are due to Miss Arisara Phuwajaroanpong and Miss Walaiporn Plirat, PhD students from the School of Medicine at WU, for their invaluable assistance in conducting the antimalarial testing. We are also grateful to the National e-Science Infrastructure Consortium (NECTEC) for the computing resources provided (www.e-science.in.th).

  1. Funding information: V.D. wishes to express gratitude to the College of Graduate Studies for the Master Excellent Scholarship (Contract No. CGS-ME 03/2020). J.J. acknowledges the support from the School of Science, University of Phayao (UP) through grants PBTSC66002 and PBTSC66023, and the Thailand Science Research and Innovation Fund in conjunction with UP (FF67-UoE).

  2. Author contributions: Vanessa Darakai: investigation, formal analysis, methodology, writing – original draft. Chuchard Punsawad: antimalarial testing resources. Jitrayut Jitonnom: software, resources, writing – review and editing. Mudtorlep Nisoa: microwave resources. Parawee Rattanakit: project administration, resources, supervision, conceptualization, writing – review and editing.

  3. Conflict of interest: Authors state no conflict of interest.

  4. Data availability statement: All data generated or analyzed during this study are included in this published article and its supplementary information files.

References

[1] Menkin-Smith L, Winders WT. Plasmodium vivax malaria. Treasure Island FL: StatPearls Publishing; 2023.Search in Google Scholar

[2] Zekar L, Sharman T. Plasmodium falciparum malaria. Treasure Island FL: StatPearls Publishing; 2023.Search in Google Scholar

[3] Monroe A, Williams NA, Ogoma S, Karema C, Okumu F. Reflections on the 2021 World Malaria Report and the future of malaria control. Malar J. 2022;1(1):154. 10.1186/s12936-022-04178-7.Search in Google Scholar PubMed PubMed Central

[4] Ippolito MM, Moser KA, Kabuya JB, Cunningham C, Juliano JJ. Antimalarial drug resistance and implications for the WHO global technical strategy. Curr Epidemiol Rep. 2021;8(2):46–62. 10.1007/s40471-021-00266-5.Search in Google Scholar PubMed PubMed Central

[5] Ansari M, Ahmed S, Abbasi A, Khan MT, Subhan M, Bukhari NA, et al. Plant mediated fabrication of silver nanoparticles, process optimization, and impact on tomato plant. Sci Rep. 2023;13(1):18048. 10.1038/s41598-023-45038-x.Search in Google Scholar PubMed PubMed Central

[6] Anis SNS, Liew WC, Mohd Marsin A, Muhamad II, Teh SH, Md Khudzari AZ. Microwave-assisted green synthesis of silver nanoparticles using pineapple leaves waste. Clean Eng Technol. 2023;15:100660. 10.1016/j.clet.2023.100660.Search in Google Scholar

[7] Khan MF, Khan MA. Plant-Derived Metal Nanoparticles (PDMNPs): Synthesis, characterization, and oxidative stress-mediated therapeutic actions. Future Pharmacol. 2023;3(1):252–95. 10.3390/futurepharmacol3010018.Search in Google Scholar

[8] Mohammadzadeh V, Barani M, Amiri MS, Taghavizadeh Yazdi ME, Hassanisaadi M, Rahdar A, et al. Applications of plant-based nanoparticles in nanomedicine: A review. Sustain Chem Pharm. 2022;25:100606. 10.1016/j.scp.2022.100606.Search in Google Scholar

[9] Bélteky P, Rónavári A, Zakupszky D, Boka E, Igaz N, Szerencsés B, et al. Are smaller nanoparticles always better? Understanding the biological effect of size-dependent silver nanoparticle aggregation under biorelevant conditions. Int J Nanomed. 2021;16:3021–40. 10.2147/IJN.S304138.Search in Google Scholar PubMed PubMed Central

[10] Ashraf H, Anjum T, Riaz S, Naseem S. Microwave-assisted green synthesis and characterization of silver nanoparticles using Melia azedarach for the management of Fusarium Wilt in tomato. Front Microbiol. 2020;11:476234. 10.3389/fmicb.2020.00238.Search in Google Scholar PubMed PubMed Central

[11] Kaur N, Singh A, Ahmad W. Microwave assisted green synthesis of silver nanoparticles and its application: A review. J Inorg Organomet Polym Mater. 2023;33(3):663–72. 10.1007/s10904-022-02470-2.Search in Google Scholar

[12] Osman MM, El-Shaheny R, Ibrahim FA. Green and fast microwave-driven synthesis of fluorescent N,S-carbon dots for determination of two 5-nitroimidazole antimicrobials. Sustain Chem Pharm. 2023;36:101255. 10.1016/j.scp.2023.101255.Search in Google Scholar

[13] Kustov L, Vikanova K. Synthesis of metal nanoparticles under microwave irradiation: Get much with less energy. Metals. 2023;13(10):1714. 10.3390/met13101714.Search in Google Scholar

[14] Goh YS, Karunakaran T, Murugaiyah V, Santhanam R, Abu Bakar MH, Ramanathan S. Accelerated solvent extractions (ASE) of Mitragyna speciosa Korth. (Kratom) leaves: Evaluation of its cytotoxicity and antinociceptive activity. Molecules. 2021;26(12):3704. 10.3390/molecules26123704.Search in Google Scholar PubMed PubMed Central

[15] Singh D, Yeou Chear NJ, Narayanan S, Leon F, Sharma A, McCurdy CR, et al. Patterns and reasons for kratom (Mitragyna speciosa) use among current and former opioid poly-drug users. J Ethnopharmacol. 2020;249:112462. 10.1016/j.jep.2019.112462.Search in Google Scholar PubMed

[16] Casey C, Conley T, Heise A, Thomas T, Ayres P. Quantitative and qualitative analysis of mitragynine in Kratom (Mitragyna speciosa) by GC–MS, LC-MS/MS and UPLC-PDA. J Regul Sci. 2015;3(2):1–14. 10.21423/JRS-V03N02P001.Search in Google Scholar

[17] Phuwajaroanpong A, Chaniad P, Plirat W, Phoopha S, Septama AW, Chukaew A, et al. Antiplasmodial properties of aqueous and ethanolic extracts of ten herbal traditional recipes used in Thailand against Plasmodium falciparum. Trop Med Infect Dis. 2022;7(12):417. 10.3390/tropicalmed7120417.Search in Google Scholar PubMed PubMed Central

[18] Kohn W, Sham LJ. Self-consistent equations including exchange and correlation effects. Phys Rev. 1965;140(4A):A1133–8. 10.1103/PhysRev.140.A1133.Search in Google Scholar

[19] Lee C, Yang W, Parr RG. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys Rev B. 1988;37(2):785–9. 10.1103/PhysRevB.37.785.Search in Google Scholar PubMed

[20] Meelua W, Wanjai T, Thinkumrob N, Oláh J, Cairns JRK, Hannongbua S, et al. A computational study of the reaction mechanism and stereospecificity of dihydropyrimidinase. Phys Chem Chem Phys. 2023;25(12):8767–78. 10.1039/D2CP05262H.Search in Google Scholar

[21] Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, et al. Gaussian 09 Rev. A.02, Wallingford, CT; 2009.Search in Google Scholar

[22] Gallegos FE, Meneses LM, Cuesta SA, Santos JC, Arias J, Carrillo P, et al. Computational modeling of the interaction of silver clusters with carbohydrates. ACS Omega. 2022;7(6):4750–6. 10.1021/acsomega.1c04149.Search in Google Scholar PubMed PubMed Central

[23] Lim JK, Kim Y, Lee SY, Joo SW. Spectroscopic analysis of l-histidine adsorbed on gold and silver nanoparticle surfaces investigated by surface-enhanced Raman scattering. Spectrochim Acta A Mol Biomol Spectrosc. 2008;69(1):286–9. 10.1016/j.saa.2007.05.007.Search in Google Scholar PubMed

[24] Abroshan H, Li G, Lin J, Kim HJ, Jin R. Molecular mechanism for the activation of Au25(SCH2CH2Ph)18 nanoclusters by imidazolium-based ionic liquids for catalysis. J Catal. 2016;337:72–9. 10.1016/j.jcat.2016.01.011.Search in Google Scholar

[25] Contreras-Torres FF. Dispersion-corrected density functional theory study of the noncovalent complexes formed with imidazo[1,2-a]pyrazines adsorbed onto silver clusters. ACS Omega. 2020;5(1):561–9. 10.1021/acsomega.9b03127.Search in Google Scholar PubMed PubMed Central

[26] Yang Y, Weaver MN, Merz Jr KM. Assessment of the “6-31 + G** + LANL2DZ” mixed basis set coupled with density functional theory methods and the effective core potential: prediction of heats of formation and ionization potentials for first-row-transition-metal complexes. J Phys Chem A. 2009;113(36):9843–51. 10.1021/jp807643p.Search in Google Scholar PubMed PubMed Central

[27] Jitonnom J, Sonntag C. Catalytic oxidation of glucose with hydrogen peroxide and colloidal gold as pseudo-homogenous catalyst: A combined experimental and theoretical investigation. Chiang Mai J Sci. 2016;43:825–33.Search in Google Scholar

[28] Jitonnom J, Meelua W. DFT study of lactide ring-opening polymerizations by aluminium trialkoxides: Understanding the effects of monomer, alkoxide substituent, solvent and metal. Chem Phys Lett. 2020;750:137482. 10.1016/j.cplett.2020.137482.Search in Google Scholar

[29] Meelua W, Wanjai T, Chokbunpiam T, Jitonnom J. Mechanism of ring-opening polymerization of l-lactide by lanthanide aryloxide: A theoretical study on the effect of the aryloxide ligands on the process. Int J Quantum Chem. 2023;123(16):e27134. 10.1002/qua.27134.Search in Google Scholar

[30] Chulasak K, Punsawad C, Rattanakit P. Silver nanoparticles synthesized from Launaea sarmentosa extract: synthesis, characterization, and antimalarial activity. Nanotechnol Env Eng. 2022;7(2):491–501. 10.1007/s41204-022-00239-z.Search in Google Scholar

[31] Firoz F, Vafaee F, Farmany A, Farhadian M, Ebrahimi S. Effect of green synthesized silver nanoparticles on optical behavior of feldspathic porcelain. Part Sci Technol. 2022;40(1):10–7. 10.1080/02726351.2021.1895385.Search in Google Scholar

[32] Huang J, Zhan G, Zheng B, Sun D, Lu F, Lin Y, et al. Biogenic silver nanoparticles by Cacumen Platycladi extract: Synthesis, formation mechanism, and antibacterial activity. Ind Eng Chem Res. 2011;50(15):9095–106. 10.1021/ie200858y.Search in Google Scholar

[33] Ahmed RH, Mustafa DE. Green synthesis of silver nanoparticles mediated by traditionally used medicinal plants in Sudan. Int Nano Lett. 2020;10(1):1–14. 10.1007/s40089-019-00291-9.Search in Google Scholar

[34] Habibullah G, Viktorova J, Ulbrich P, Ruml T. Effect of the physicochemical changes in the antimicrobial durability of green synthesized silver nanoparticles during their long-term storage. RSC Adv. 2022;12(47):30386–403. 10.1039/D2RA04667A.Search in Google Scholar

[35] Ali K, Ahmed B, Dwivedi S, Saquib Q, Al-Khedhairy AA, Musarrat J. Microwave accelerated green synthesis of stable silver nanoparticles with Eucalyptus globulus leaf extract and their antibacterial and antibiofilm activity on clinical isolates. PLoS One. 2015;10(7):e0131178. 10.1371/journal.pone.0131178.Search in Google Scholar PubMed PubMed Central

[36] Saha N, Trivedi P, Dutta Gupta S. Surface plasmon resonance (SPR) based optimization of biosynthesis of silver nanoparticles from Rhizome extract of Curculigo orchioides Gaertn. and its antioxidant potential. J Clust Sci. 2016;27:1893–912.10.1007/s10876-016-1050-7Search in Google Scholar

[37] Amendola V, Bakr OM, Stellacci F. A Study of the surface plasmon resonance of silver nanoparticles by the discrete dipole approximation method: Effect of shape, size, structure, and assembly. Plasmonics. 2010;5(1):85–97. 10.1007/s11468-009-9120-4.Search in Google Scholar

[38] Ayinde WB, Gitari WM, Samie A. Optimization of microwave-assisted synthesis of silver nanoparticle by Citrus paradisi peel and its application against pathogenic water strain. Green Chem Lett Rev. 2019;12(3):225–34. 10.1080/17518253.2019.1627427.Search in Google Scholar

[39] Barani H, Mahltig B. Microwave-assisted synthesis of silver nanoparticles: Effect of reaction temperature and precursor concentration on fluorescent property. J Clust Sci. 2022;33(1):101–11. 10.1007/s10876-020-01945-x.Search in Google Scholar

[40] Palma V, Barba D, Cortese M, Martino M, Renda S, Meloni E. Microwaves and heterogeneous catalysis: A teview on selected catalytic processes. Catalysts. 2020;10(2):246. 10.3390/catal10020246.Search in Google Scholar

[41] Sweygers N, Kamali M, Aminabhavi TM, Dewil R, Appels L. Efficient microwave-assisted production of furanics and hydrochar from bamboo (Phyllostachys nigra “Boryana”) in a biphasic reaction system: effect of inorganic salts. Biomass Convers Bior. 2022;12(1):173–81. 10.1007/s13399-021-01372-6.Search in Google Scholar

[42] Khalil M, Ismail E, El-Baghdady K, Mohamed D. Green synthesis of silver nanoparticles using olive leaf extract and its antibacterial activity. Arab J Chem. 2013;7:1131–9. 10.1016/j.arabjc.2013.04.007.Search in Google Scholar

[43] Viorica R, Pawel P, Płociński T, Gloc M, Dobrucka R, Kurzydłowski KJ, et al. Consideration of a new approach to clarify the mechanism formation of AgNPs, AgNCl and AgNPs@AgNCl synthesized by biological method. Discov Nano. 2023;18(1):2. 10.1186/s11671-023-03777-w.Search in Google Scholar PubMed PubMed Central

[44] Jain N, Jain P, Rajput D, Patil UK. Green synthesized plant-based silver nanoparticles: therapeutic prospective for anticancer and antiviral activity. Micro Nano Syst Lett. 2021;9(1):5. 10.1186/s40486-021-00131-6.Search in Google Scholar

[45] Khwannimit D, Maungchang R, Rattanakit P. Green synthesis of silver nanoparticles using Clitoria ternatea flower: an efficient catalyst for removal of methyl orange. Int J Env Anal Chem. 2022;102(17):5247–63. 10.1080/03067319.2020.1793974.Search in Google Scholar

[46] Odira HO, Mitema SO, Mapenay IM, Moriasi GA. Anti-inflammatory, analgesic, and cytotoxic effects of the phytexponent: A polyherbal formulation. J Evid Based Integr Med. 2022;27:2515690x221082986. 10.1177/2515690x221082986.Search in Google Scholar PubMed PubMed Central

[47] Indrayanto G, Putra GS, Suhud F. Validation of in-vitro bioassay methods: Application in herbal drug research. Profiles Drug Subst Excip Relat Methodol. 2021;46:273–307. 10.1016/bs.podrm.2020.07.005.Search in Google Scholar PubMed

[48] Awouafack MD, McGaw LJ, Gottfried S, Mbouangouere R, Tane P, Spiteller M, et al. Antimicrobial activity and cytotoxicity of the ethanol extract, fractions and eight compounds isolated from Eriosema robustum (Fabaceae). BMC Complement Altern Med. 2013;13:289. 10.1186/1472-6882-13-289.Search in Google Scholar PubMed PubMed Central

[49] Jiang C, Jiang Z, Zhu S, Amulraj J, Deenadayalan V, Jacob JA, et al. Biosynthesis of silver nanoparticles and the identification of possible reductants for the assessment of in vitro cytotoxic and in vivo antitumor effects. J Drug Deliv Sci Technol. 2021;63:102444. 10.1016/j.jddst.2021.102444.Search in Google Scholar

[50] Yuniarti R, Nadia S, Alamanda A, Zubir M, Syahputra RA, Nizam M. Characterization, phytochemical screenings and antioxidant activity test of Kratom leaf ethanol extract (Mitragyna speciosa Korth) using DPPH method. J Phys Conf Ser. 2020;1462(1):012026. 10.1088/1742-6596/1462/1/012026.Search in Google Scholar

[51] Luna-Vázquez-Gómez R, Arellano-García ME, García-Ramos JC, Radilla-Chávez P, Salas-Vargas DS, Casillas-Figueroa F, et al. Hemolysis of human erythrocytes by Argovit™ AgNPs from healthy and diabetic donors: An in vitro study. Mater (Basel). 2021;14(11):2792. 10.3390/ma14112792.Search in Google Scholar PubMed PubMed Central

[52] Plowe CV. Malaria chemoprevention and drug resistance: a review of the literature and policy implications. Malar J. 2022;21(1):104. 10.1186/s12936-022-04115-8.Search in Google Scholar PubMed PubMed Central

[53] Desta K, Amare M. Validated UV-Visible spectrometry using water as a solvent for determination of chloroquine in tablet samples. Chem Int. 2017;3(3):288–95.Search in Google Scholar

[54] Chaves JB, Portugal Tavares de Moraes B, Regina Ferrarini S, Noé da Fonseca F, Silva AR, Gonçalves-de-Albuquerque CF. Potential of nanoformulations in malaria treatment. Front Pharmacol. 2022;13:999300. 10.3389/fphar.2022.999300.Search in Google Scholar PubMed PubMed Central

[55] Nie P, Zhao Y, Xu H. Synthesis, applications, toxicity and toxicity mechanisms of silver nanoparticles: A review. Ecotoxicol Env Saf. 2023;253:114636. 10.1016/j.ecoenv.2023.114636.Search in Google Scholar PubMed

Received: 2023-12-18
Accepted: 2024-03-07
Published Online: 2024-04-04

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Green polymer electrolyte and activated charcoal-based supercapacitor for energy harvesting application: Electrochemical characteristics
  3. Research on the adsorption of Co2+ ions using halloysite clay and the ability to recover them by electrodeposition method
  4. Simultaneous estimation of ibuprofen, caffeine, and paracetamol in commercial products using a green reverse-phase HPTLC method
  5. Isolation, screening and optimization of alkaliphilic cellulolytic fungi for production of cellulase
  6. Functionalized gold nanoparticles coated with bacterial alginate and their antibacterial and anticancer activities
  7. Comparative analysis of bio-based amino acid surfactants obtained via Diels–Alder reaction of cyclic anhydrides
  8. Biosynthesis of silver nanoparticles on yellow phosphorus slag and its application in organic coatings
  9. Exploring antioxidant potential and phenolic compound extraction from Vitis vinifera L. using ultrasound-assisted extraction
  10. Manganese and copper-coated nickel oxide nanoparticles synthesized from Carica papaya leaf extract induce antimicrobial activity and breast cancer cell death by triggering mitochondrial caspases and p53
  11. Insight into heating method and Mozafari method as green processing techniques for the synthesis of micro- and nano-drug carriers
  12. Silicotungstic acid supported on Bi-based MOF-derived metal oxide for photodegradation of organic dyes
  13. Synthesis and characterization of capsaicin nanoparticles: An attempt to enhance its bioavailability and pharmacological actions
  14. Synthesis of Lawsonia inermis-encased silver–copper bimetallic nanoparticles with antioxidant, antibacterial, and cytotoxic activity
  15. Facile, polyherbal drug-mediated green synthesis of CuO nanoparticles and their potent biological applications
  16. Zinc oxide-manganese oxide/carboxymethyl cellulose-folic acid-sesamol hybrid nanomaterials: A molecularly targeted strategy for advanced triple-negative breast cancer therapy
  17. Exploring the antimicrobial potential of biogenically synthesized graphene oxide nanoparticles against targeted bacterial and fungal pathogens
  18. Biofabrication of silver nanoparticles using Uncaria tomentosa L.: Insight into characterization, antibacterial activities combined with antibiotics, and effect on Triticum aestivum germination
  19. Membrane distillation of synthetic urine for use in space structural habitat systems
  20. Investigation on mechanical properties of the green synthesis bamboo fiber/eggshell/coconut shell powder-based hybrid biocomposites under NaOH conditions
  21. Green synthesis of magnesium oxide nanoparticles using endophytic fungal strain to improve the growth, metabolic activities, yield traits, and phenolic compounds content of Nigella sativa L.
  22. Estimation of greenhouse gas emissions from rice and annual upland crops in Red River Delta of Vietnam using the denitrification–decomposition model
  23. Synthesis of humic acid with the obtaining of potassium humate based on coal waste from the Lenger deposit, Kazakhstan
  24. Ascorbic acid-mediated selenium nanoparticles as potential antihyperuricemic, antioxidant, anticoagulant, and thrombolytic agents
  25. Green synthesis of silver nanoparticles using Illicium verum extract: Optimization and characterization for biomedical applications
  26. Antibacterial and dynamical behaviour of silicon nanoparticles influenced sustainable waste flax fibre-reinforced epoxy composite for biomedical application
  27. Optimising coagulation/flocculation using response surface methodology and application of floc in biofertilisation
  28. Green synthesis and multifaceted characterization of iron oxide nanoparticles derived from Senna bicapsularis for enhanced in vitro and in vivo biological investigation
  29. Potent antibacterial nanocomposites from okra mucilage/chitosan/silver nanoparticles for multidrug-resistant Salmonella Typhimurium eradication
  30. Trachyspermum copticum aqueous seed extract-derived silver nanoparticles: Exploration of their structural characterization and comparative antibacterial performance against gram-positive and gram-negative bacteria
  31. Microwave-assisted ultrafine silver nanoparticle synthesis using Mitragyna speciosa for antimalarial applications
  32. Green synthesis and characterisation of spherical structure Ag/Fe2O3/TiO2 nanocomposite using acacia in the presence of neem and tulsi oils
  33. Green quantitative methods for linagliptin and empagliflozin in dosage forms
  34. Enhancement efficacy of omeprazole by conjugation with silver nanoparticles as a urease inhibitor
  35. Residual, sequential extraction, and ecological risk assessment of some metals in ash from municipal solid waste incineration, Vietnam
  36. Green synthesis of ZnO nanoparticles using the mangosteen (Garcinia mangostana L.) leaf extract: Comparative preliminary in vitro antibacterial study
  37. Simultaneous determination of lesinurad and febuxostat in commercial fixed-dose combinations using a greener normal-phase HPTLC method
  38. A greener RP-HPLC method for quaternary estimation of caffeine, paracetamol, levocetirizine, and phenylephrine acquiring AQbD with stability studies
  39. Optimization of biomass durian peel as a heterogeneous catalyst in biodiesel production using microwave irradiation
  40. Thermal treatment impact on the evolution of active phases in layered double hydroxide-based ZnCr photocatalysts: Photodegradation and antibacterial performance
  41. Preparation of silymarin-loaded zein polysaccharide core–shell nanostructures and evaluation of their biological potentials
  42. Preparation and characterization of composite-modified PA6 fiber for spectral heating and heat storage applications
  43. Preparation and electrocatalytic oxygen evolution of bimetallic phosphates (NiFe)2P/NF
  44. Rod-shaped Mo(vi) trichalcogenide–Mo(vi) oxide decorated on poly(1-H pyrrole) as a promising nanocomposite photoelectrode for green hydrogen generation from sewage water with high efficiency
  45. Green synthesis and studies on citrus medica leaf extract-mediated Au–ZnO nanocomposites: A sustainable approach for efficient photocatalytic degradation of rhodamine B dye in aqueous media
  46. Cellulosic materials for the removal of ciprofloxacin from aqueous environments
  47. The analytical assessment of metal contamination in industrial soils of Saudi Arabia using the inductively coupled plasma technology
  48. The effect of modified oily sludge on the slurry ability and combustion performance of coal water slurry
  49. Eggshell waste transformation to calcium chloride anhydride as food-grade additive and eggshell membranes as enzyme immobilization carrier
  50. Synthesis of EPAN and applications in the encapsulation of potassium humate
  51. Biosynthesis and characterization of silver nanoparticles from Cedrela toona leaf extracts: An exploration into their antibacterial, anticancer, and antioxidant potential
  52. Enhancing mechanical and rheological properties of HDPE films through annealing for eco-friendly agricultural applications
  53. Immobilisation of catalase purified from mushroom (Hydnum repandum) onto glutaraldehyde-activated chitosan and characterisation: Its application for the removal of hydrogen peroxide from artificial wastewater
  54. Sodium titanium oxide/zinc oxide (STO/ZnO) photocomposites for efficient dye degradation applications
  55. Effect of ex situ, eco-friendly ZnONPs incorporating green synthesised Moringa oleifera leaf extract in enhancing biochemical and molecular aspects of Vicia faba L. under salt stress
  56. Biosynthesis and characterization of selenium and silver nanoparticles using Trichoderma viride filtrate and their impact on Culex pipiens
  57. Photocatalytic degradation of organic dyes and biological potentials of biogenic zinc oxide nanoparticles synthesized using the polar extract of Cyperus scariosus R.Br. (Cyperaceae)
  58. Assessment of antiproliferative activity of green-synthesized nickel oxide nanoparticles against glioblastoma cells using Terminalia chebula
  59. Chlorine-free synthesis of phosphinic derivatives by change in the P-function
  60. Anticancer, antioxidant, and antimicrobial activities of nanoemulsions based on water-in-olive oil and loaded on biogenic silver nanoparticles
  61. Study and mechanism of formation of phosphorus production waste in Kazakhstan
  62. Synthesis and stabilization of anatase form of biomimetic TiO2 nanoparticles for enhancing anti-tumor potential
  63. Microwave-supported one-pot reaction for the synthesis of 5-alkyl/arylidene-2-(morpholin/thiomorpholin-4-yl)-1,3-thiazol-4(5H)-one derivatives over MgO solid base
  64. Screening the phytochemicals in Perilla leaves and phytosynthesis of bioactive silver nanoparticles for potential antioxidant and wound-healing application
  65. Graphene oxide/chitosan/manganese/folic acid-brucine functionalized nanocomposites show anticancer activity against liver cancer cells
  66. Nature of serpentinite interactions with low-concentration sulfuric acid solutions
  67. Multi-objective statistical optimisation utilising response surface methodology to predict engine performance using biofuels from waste plastic oil in CRDi engines
  68. Microwave-assisted extraction of acetosolv lignin from sugarcane bagasse and electrospinning of lignin/PEO nanofibres for carbon fibre production
  69. Biosynthesis, characterization, and investigation of cytotoxic activities of selenium nanoparticles utilizing Limosilactobacillus fermentum
  70. Highly photocatalytic materials based on the decoration of poly(O-chloroaniline) with molybdenum trichalcogenide oxide for green hydrogen generation from Red Sea water
  71. Highly efficient oil–water separation using superhydrophobic cellulose aerogels derived from corn straw
  72. Beta-cyclodextrin–Phyllanthus emblica emulsion for zinc oxide nanoparticles: Characteristics and photocatalysis
  73. Assessment of antimicrobial activity and methyl orange dye removal by Klebsiella pneumoniae-mediated silver nanoparticles
  74. Influential eradication of resistant Salmonella Typhimurium using bioactive nanocomposites from chitosan and radish seed-synthesized nanoselenium
  75. Antimicrobial activities and neuroprotective potential for Alzheimer’s disease of pure, Mn, Co, and Al-doped ZnO ultra-small nanoparticles
  76. Green synthesis of silver nanoparticles from Bauhinia variegata and their biological applications
  77. Synthesis and optimization of long-chain fatty acids via the oxidation of long-chain fatty alcohols
  78. Eminent Red Sea water hydrogen generation via a Pb(ii)-iodide/poly(1H-pyrrole) nanocomposite photocathode
  79. Green synthesis and effective genistein production by fungal β-glucosidase immobilized on Al2O3 nanocrystals synthesized in Cajanus cajan L. (Millsp.) leaf extracts
  80. Green stability-indicating RP-HPTLC technique for determining croconazole hydrochloride
  81. Green synthesis of La2O3–LaPO4 nanocomposites using Charybdis natator for DNA binding, cytotoxic, catalytic, and luminescence applications
  82. Eco-friendly drugs induce cellular changes in colistin-resistant bacteria
  83. Tangerine fruit peel extract mediated biogenic synthesized silver nanoparticles and their potential antimicrobial, antioxidant, and cytotoxic assessments
  84. Green synthesis on performance characteristics of a direct injection diesel engine using sandbox seed oil
  85. A highly sensitive β-AKBA-Ag-based fluorescent “turn off” chemosensor for rapid detection of abamectin in tomatoes
  86. Green synthesis and physical characterization of zinc oxide nanoparticles (ZnO NPs) derived from the methanol extract of Euphorbia dracunculoides Lam. (Euphorbiaceae) with enhanced biosafe applications
  87. Detection of morphine and data processing using surface plasmon resonance imaging sensor
  88. Effects of nanoparticles on the anaerobic digestion properties of sulfamethoxazole-containing chicken manure and analysis of bio-enzymes
  89. Bromic acid-thiourea synergistic leaching of sulfide gold ore
  90. Green chemistry approach to synthesize titanium dioxide nanoparticles using Fagonia Cretica extract, novel strategy for developing antimicrobial and antidiabetic therapies
  91. Green synthesis and effective utilization of biogenic Al2O3-nanocoupled fungal lipase in the resolution of active homochiral 2-octanol and its immobilization via aluminium oxide nanoparticles
  92. Eco-friendly RP-HPLC approach for simultaneously estimating the promising combination of pentoxifylline and simvastatin in therapeutic potential for breast cancer: Appraisal of greenness, whiteness, and Box–Behnken design
  93. Use of a humidity adsorbent derived from cockleshell waste in Thai fried fish crackers (Keropok)
  94. One-pot green synthesis, biological evaluation, and in silico study of pyrazole derivatives obtained from chalcones
  95. Bio-sorption of methylene blue and production of biofuel by brown alga Cystoseira sp. collected from Neom region, Kingdom of Saudi Arabia
  96. Synthesis of motexafin gadolinium: A promising radiosensitizer and imaging agent for cancer therapy
  97. The impact of varying sizes of silver nanoparticles on the induction of cellular damage in Klebsiella pneumoniae involving diverse mechanisms
  98. Microwave-assisted green synthesis, characterization, and in vitro antibacterial activity of NiO nanoparticles obtained from lemon peel extract
  99. Rhus microphylla-mediated biosynthesis of copper oxide nanoparticles for enhanced antibacterial and antibiofilm efficacy
  100. Harnessing trichalcogenide–molybdenum(vi) sulfide and molybdenum(vi) oxide within poly(1-amino-2-mercaptobenzene) frameworks as a photocathode for sustainable green hydrogen production from seawater without sacrificial agents
  101. Magnetically recyclable Fe3O4@SiO2 supported phosphonium ionic liquids for efficient and sustainable transformation of CO2 into oxazolidinones
  102. A comparative study of Fagonia arabica fabricated silver sulfide nanoparticles (Ag2S) and silver nanoparticles (AgNPs) with distinct antimicrobial, anticancer, and antioxidant properties
  103. Visible light photocatalytic degradation and biological activities of Aegle marmelos-mediated cerium oxide nanoparticles
  104. Physical intrinsic characteristics of spheroidal particles in coal gasification fine slag
  105. Exploring the effect of tea dust magnetic biochar on agricultural crops grown in polycyclic aromatic hydrocarbon contaminated soil
  106. Crosslinked chitosan-modified ultrafiltration membranes for efficient surface water treatment and enhanced anti-fouling performances
  107. Study on adsorption characteristics of biochars and their modified biochars for removal of organic dyes from aqueous solution
  108. Zein polymer nanocarrier for Ocimum basilicum var. purpurascens extract: Potential biomedical use
  109. Green synthesis, characterization, and in vitro and in vivo biological screening of iron oxide nanoparticles (Fe3O4) generated with hydroalcoholic extract of aerial parts of Euphorbia milii
  110. Novel microwave-based green approach for the synthesis of dual-loaded cyclodextrin nanosponges: Characterization, pharmacodynamics, and pharmacokinetics evaluation
  111. Bi2O3–BiOCl/poly-m-methyl aniline nanocomposite thin film for broad-spectrum light-sensing
  112. Green synthesis and characterization of CuO/ZnO nanocomposite using Musa acuminata leaf extract for cytotoxic studies on colorectal cancer cells (HCC2998)
  113. Review Articles
  114. Materials-based drug delivery approaches: Recent advances and future perspectives
  115. A review of thermal treatment for bamboo and its composites
  116. An overview of the role of nanoherbicides in tackling challenges of weed management in wheat: A novel approach
  117. An updated review on carbon nanomaterials: Types, synthesis, functionalization and applications, degradation and toxicity
  118. Special Issue: Emerging green nanomaterials for sustainable waste management and biomedical applications
  119. Green synthesis of silver nanoparticles using mature-pseudostem extracts of Alpinia nigra and their bioactivities
  120. Special Issue: New insights into nanopythotechnology: current trends and future prospects
  121. Green synthesis of FeO nanoparticles from coffee and its application for antibacterial, antifungal, and anti-oxidation activity
  122. Dye degradation activity of biogenically synthesized Cu/Fe/Ag trimetallic nanoparticles
  123. Special Issue: Composites and green composites
  124. Recent trends and advancements in the utilization of green composites and polymeric nanocarriers for enhancing food quality and sustainable processing
  125. Retraction
  126. Retraction of “Biosynthesis and characterization of silver nanoparticles from Cedrela toona leaf extracts: An exploration into their antibacterial, anticancer, and antioxidant potential”
  127. Retraction of “Photocatalytic degradation of organic dyes and biological potentials of biogenic zinc oxide nanoparticles synthesized using the polar extract of Cyperus scariosus R.Br. (Cyperaceae)”
  128. Retraction to “Green synthesis on performance characteristics of a direct injection diesel engine using sandbox seed oil”
Downloaded on 22.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/gps-2023-0257/html
Scroll to top button