Next Article in Journal
Comparison of the Rate Constants of OH, SO4•−, CO3•−, Cl2•−, Cl, ClO and H Reactions with Organic Water Contaminants
Previous Article in Journal
DFT Study of Functionalized Benzoxazole-Based D–π–A Architectures: Influence of Ionic Fragments on Optical Properties and Their Potential in OLED and Solar Cell Devices
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Green Synthesized Silver Nanoparticles from Biowaste for Rapid Dye Degradation: Experimental Investigation and Computational Mechanistic Insights

by
Tanakorn Wonglakhon
1,2,
Areeya Chonsakon
3,4,
Prawit Nuengmatcha
3,4,
Benjawan Ninwong
3,4,
Dirk Zahn
5 and
Yanisa Thepchuay
3,4,6,*
1
Futuristic Science Research Center, School of Science, Walailak University, Nakhon Si Thammarat 80160, Thailand
2
Research Center for Theoretical Simulation and Applied Research in Bioscience and Sensing, Walailak University, Nakhon Si Thammarat 80160, Thailand
3
Center of Excellence in Nanomaterials Chemistry, Faculty of Science and Technology, Nakhon Si Thammarat Rajabhat University, Nakhon Si Thammarat 80280, Thailand
4
Department of Chemistry, Faculty of Science and Technology, Nakhon Si Thammarat Rajabhat University, Nakhon Si Thammarat 80280, Thailand
5
Lehrstuhl für Theoretische Chemie/Computer Chemie Centrum, Friedrich-Alexander-Universität Erlangen-Nürnberg, Nägelsbachstraße 25, 91052 Erlangen, Germany
6
Flow Innovation-Research for Science and Technology Laboratories (Firstlabs), Bangkok 10400, Thailand
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(18), 3738; https://doi.org/10.3390/molecules30183738
Submission received: 12 August 2025 / Revised: 7 September 2025 / Accepted: 9 September 2025 / Published: 15 September 2025
(This article belongs to the Special Issue Nano and Micro Materials in Green Chemistry)

Abstract

Silver nanoparticles (Ag NPs) green-synthesized using Nypa fruticans fruit husk (NF) extract were applied as catalysts for the degradation of organic dyes in water for the first time. The synthesized Ag NPs, which were well-dispersed, highly stable, and small in size with an average diameter of ~4 nm, efficiently catalyzed the degradation of methyl orange (MO) in the presence of NaBH 4 , achieving complete degradation (>99%) within one minute under optimized conditions. The application to a commercial synthetic dye resulted in over 89% degradation within five minutes. To elucidate the degradation mechanism at the atomistic level, molecular dynamics (MD) simulations and density functional theory (DFT) calculations were employed. MD simulations revealed the adsorption behavior of MO on the Ag(111) surface. DFT calculations clarified the reaction pathway of MO degradation, identifying direct hydride transfer from BH 4 to the azo group of MO as the rate-determining step, with the subsequent step influenced by the pH conditions. These findings illustrate the potential of NF extract in the green synthesis of catalytically active Ag NPs and contribute to understanding their role in dye degradation processes relevant to environmental remediation.

Graphical Abstract

1. Introduction

Water pollution from dyes poses severe risks to both ecosystems and human health. Industrial and domestic activities contribute significantly to the release of these contaminants, which disrupt surface water chemistry and harm aquatic life. Annually, approximately 7 × 107 tons of synthetic dyes are produced worldwide, with textile industries consuming over 10,000 tons [1]. The textile industry discharges large volumes of highly colored wastewater containing various synthetic dyes, including azo, reactive, and sulfide dyes [2]. However, even trace amounts of these dyes (e.g., 1 × 10−3 mg/L) are hazardous and can cause mutagenic and carcinogenic effects that impact the human nervous, digestive, renal, and hepatic systems [3].
In essence, effective degradation or removal of dyes is essential and urgently needed. Methods for detecting and removing organic pollutants are typically divided into chemical (e.g., reduction, photochemical degradation), physical (e.g., adsorption, filtration), and biological processes, with adsorption being a simple and commonly used approach [4,5]. In response to the limitations of traditional techniques, researchers are increasingly focusing on simpler and more cost-effective alternatives. One promising option is a colorimetric method using biologically synthesized nanoparticles, particularly silver nanoparticles (Ag NPs). These nanoparticles show potential for catalytically reducing dyes due to their unique properties, such as high surface area, strong catalytic activity, and the ability to change color when interacting with pollutants [6,7,8,9,10,11]. Furthermore, Ag NP-based methods are relatively easy to implement and less costly than more complex and expensive instruments, rendering Ag NPs one of the practical choices for environmental applications.
Ag NPs synthesized from various natural extracts, such as those from mango leaf [6], cauliflower waste [8], and Epilobium parviflorum green tea [7], have been shown to efficiently degrade hazardous organic dyes, and even detect heavy metal ions. In our recent study [12], Ag NPs were successfully synthesized using the Nypa fruticans fruit husk (NF) extract [13,14], which is a common palm species grown in Thailand, particularly in the south. Nipa palms have been used for various products such as syrup, sugar, and vinegar derived from the palm sap [15], leaving large quantities of nipa palm waste unemployed. This has inspired us to transform waste into valuable resources by exploring its potential for Ag NP synthesis. This biowaste-based synthesis adds value to waste materials and supports sustainable waste management by producing valuable nanomaterials, in line with the Sustainable Development Goals (SDGs) [16]. Our synthesized Ag NPs have already been shown to be highly selective toward Fe2+ metal ions [12].
While Ag NPs are known to act as electron relays that facilitate color changes and catalytic activity, in the present study, we present a combined experimental and theoretical analysis of the degradation of hazardous methyl orange (MO) dye. We thus extend the application perspectives of the Ag NPs synthesized using the NF extract, for the first time, to the efficient and eco-friendly degradation of organic dyes and provide mechanistic insights from molecular dynamics simulations and quantum chemical calculations, respectively.

2. Results and Discussion

2.1. Optimized Conditions for the Synthesis of Ag NPs

In our previous study [12], we optimized the synthesis conditions of Ag NPs using NF extract by analyzing their surface plasmon resonance (SPR) bands at 400–425 nm [17] via UV–Vis spectroscopy. The optimal conditions were determined through systematic variations in the key reaction parameters: pH, Ag NO 3 concentration, NF extract concentration, volume ratio of NF extract to Ag NO 3 , temperature, and reaction time. The highest SPR intensity with a narrow band, indicating well-dispersed Ag NPs, was achieved at pH 9.0. The optimal Ag NO 3 concentration was 1.0 mM, while 0.2% (w/v) NF extract was sufficient for complete reduction and stabilization. A volume ratio of NF extract to Ag NO 3 of 1:20 provided maximum yield. Temperature optimization revealed that 60 °C was appropriate, as higher temperatures led to precipitation. Lastly, a reaction time of 45 min was found to be optimal, as prolonged durations led to particle aggregation. These optimized conditions (Table 1) were established to ensure the formation of stable, uniformly distributed Ag NPs with controlled size and shape.

2.2. Characteristics and Stability of the Synthesized Ag NPs

As reported in our previous study [12], we established that NF extract-mediated Ag NPs exhibit stable, well-dispersed, highly crystalline structures with a mean size of ∼4 nm (Figure 1a–d). Elemental analysis via EDX confirmed the presence of Ag along with trace elements from the extract (Figure 1e). FT-IR spectra indicated strong interactions between phytochemicals and the Ag NP surface, particularly through O–H and aromatic groups (Figure 1f). The stability test showed that our prepared Ag NPs were stable for at least two weeks, showing potential for long-term storage. Theoretical studies using DFT and MD simulations revealed that phenolic compounds play a key role as reducing and stabilizing agents, showing a high affinity for the Ag(111) surface, which prevents uncontrolled aggregation. At the optimal pH of 9, deprotonation of phenolic groups resulted in negatively charged, phenolic compound-functionalized Ag NPs, as evidenced by a zeta potential of −41.9 mV, which contributed to enhanced colloidal stability by preventing aggregation. Building on these findings, the present study extends the application of the Ag NPs synthesized using the same synthesis conditions to the degradation of methyl orange dye, which is a toxic and non-biodegradable azo dye, as illustrated in subsequent sections. To confirm batch reproducibility, the newly synthesized Ag NPs were re-characterized by UV–Vis spectrophotometry. The spectrum showed a characteristic SPR band at 424 nm with a peak absorbance of 1.6415, closely matching our previous batch (1.6493) [12]; the peak absorbances differ by less than 0.5%. Figure S1 displays the raw spectral overlay: the plasmon peak position and the peak profile are essentially identical; any differences are predominantly on the long-wavelength tail (λ > 450 nm). These results verify that the Ag NPs used here were distinct yet reproducible under the same synthesis conditions.

2.3. Catalytic Reduction of Methyl Orange Dye

To illustrate the application of our synthesized Ag NPs, we assess the function of the synthesized Ag NPs to act as the catalysts for the dye degradation. For this, the catalytic activity of the synthesized Ag NPs for degrading MO was evaluated in the presence of NaBH 4 . In aqueous medium, the azo (–N=N–) group of the MO shows a characteristic absorption peak at 464 nm. Therefore, we focus on the spectrum change at this wavelength, which shows the maximum absorbance at 464 nm, to track the dye degradation.
Firstly, the MO dyes were found to degrade very slowly in the absence of Ag NPs as indicated by the very slow decrease in the absorbance spectrum as a function of time (Figure 2). This may be hindered by the electrostatic repulsion between the negatively charged borohydride ion ( BH 4 ) and the negatively charged MO dye (an anionic dye), making the reduction reaction difficult to surpass the energy barrier. Upon the introduction of the Ag NPs to the system, the MO degradation is expected to proceed faster since the Ag NPs act as the electron relay to facilitate the NaBH 4 to reduce the MO dyes.
To find the suitable concentration of NaBH 4 for the MO degradation, we studied six concentrations of 0.5 mL NaBH 4 (i.e., 0.05, 0.10, 0.15, 0.20, 0.25, and 0.30 M) while keeping the Ag NP solution at 50 μL and the MO solution at 2.5 mL (0.1 mM). The catalytic efficiency was then assessed as the percent of MO degradation as a function of time. Figure 3a shows the initial MO solution and the resulting decolorized solution after degradation. As shown in Figure 3d,e, at a given time, the degradation percentages of MO increased significantly with the increase in the NaBH 4 concentration. However, upon increasing the concentration of NaBH 4 to 0.2 M, the MO was effectively degraded over 99%, which is not significantly different when the NaBH 4 concentration was increased to 0.25 M (Figure 3e). This corresponds to a rapid decrease in the absorbance spectrum within 2.5 min at 0.20 M NaBH 4 (Figure 3c vs. Figure 3b). In addition, when the NaBH 4 concentration increases to 0.3 M, we observed a decrease in the degradation percentage and the rate constant. Therefore, 0.20 M NaBH 4 is already enough to reduce the MO dyes on the Ag NPs. As shown in Figure 3f, the MO degradation follows the pseudo-first-order kinetics, ln(At/A0) = −kt, where k is the degradation rate constant and A0 and At are the absorbances at t = 0 and time t, respectively. From the plot of ln(At/A0) as a function of time (Figure 3f), a rate constant for each NaBH 4 concentration was derived and is shown in Table 2. It can be concluded that the rate constant increased with the NaBH 4 concentration up to 0.25 M but declined at 0.30 M. Because >99% MO degradation was achieved within 2.5 min at 0.20 M, we selected 0.20 M NaBH 4 as the working concentration.
The effect of the amount of Ag NP catalyst on the catalytic reduction of MO azo dye was further investigated by varying the volume of the Ag NP catalyst from 25 μL to 100 μL at a fixed concentration of NaBH 4 (0.2 M) and MO dye (0.1 mM). The reduction process was then monitored by the decrease in the absorbance peak at 465 nm and simultaneous increase in the absorbance at 250 nm (Figure 4). In analogy to the above study, the pseudo-first-order kinetics was used to assess the catalytic performance of Ag NPs. When the Ag NP volume increases from 25 to 100 μL, the rate constant was found to increase from 0.523 to 2.763 min−1 while the time for the complete MO degradation decreased from 4.5 to just 1 min (Table 3). At 50 μL, the MO degradation percentage already reached 95%. To ensure economical application of Ag NPs, we therefore chose 50 μL Ag NPs.

2.4. Comparison with Other Studies

A comparison with other studies on using green-synthesized Ag NPs to degrade the MO dye in the presence of NaBH 4 (from 2014 to the present) is summarized in Table 4. For a fair kinetic comparison, Table 4 excludes non-green syntheses and non-Ag NP catalysts for dye degradation (e.g., peroxymonosulfate (PMS)/Fenton-like oxidation [18,19]). As can be observed, the reported kinetic parameters vary considerably depending on factors such as the source of nanoparticle synthesis and catalyst size and morphology, including the Ag NP and NaBH 4 concentrations. Compared to these studies, our Ag NPs synthesized using NF extract achieve more than 99% MO degradation within one minute and exhibit among the higher rate constant and shorter reduction time reported. This superior catalytic performance may be attributed to the small particle size and good dispersion of the Ag NPs in solution, which facilitate the interactions between MO dye molecules and NaBH 4 . Encouraged by these promising results, we extended the application of our Ag NPs to the degradation of a commercially available synthetic dye obtained from a local market in Thailand, as described in the following section.

2.5. Catalytic Reduction of Commercial Synthetic Dye for Cotton Fabrics

To demonstrate the effectiveness of our synthesized Ag NPs, we investigated the catalytic reduction of a commercial synthetic dye, with direct deep yellow. Although this dye lacks a unique identification code or specific chemical name, it mainly contains azo, phenyl, and sulfonic acid groups and is widely used for coloring cotton fabrics in Thailand. The experiment was carried out at room temperature by mixing 2.5 mL of 1000 ppm synthetic dye (prepared with deionized water) and 0.5 mL of 0.2 M NaBH 4 with 50 μL of Ag NP solution. At the beginning, the color solution appeared a deep yellow color. As the reaction progressed, the deep yellow solution decolorized, accompanied by a rapid decrease in the absorbance at 465 nm (Figure 5a), yielding more than 89% degradation within only 5 min (Figure 5b). In real water (non-potable tap water), degradation was ~83% at 5 min and reached ~86% by 6 min for two independent samples (Figure S2), consistent with a modest matrix effect, while maintaining high catalytic performance. Therefore, these results indicate that Ag NPs synthesized using NF extract are promising candidates for the catalytic treatment of wastewater containing non-biodegradable azo dye.

2.6. Insights into the Degradation Mechanism of Methyl Orange via Molecular Simulations

2.6.1. MO Dye Adsorption Mechanism via MD Simulations

While several studies have explored the use of Ag NPs in the degradation of MO dye (as shown in Table 4), they commonly suggest that Ag NPs act as electron relays that facilitate the degradation process. However, the underlying mechanisms—particularly at the atomistic level—remain poorly understood. From the molecular dynamics perspective, which is well suited for investigating adsorption behavior, no prior simulations have specifically examined the adsorption of MO onto silver nanoparticle surfaces.
On other materials, a few theoretical studies have employed MD simulations to investigate the adsorption behavior of MO dye on solid surfaces. For example, Fahimirad et al. conducted combined experimental and theoretical work on MO photodegradation using Au-ZnO catalysts [56]. Their theoretical study employed the Monte Carlo method, which is similar in concept to MD, to model MO adsorption on Au-ZnO(111) with 10 water molecules. Despite the arbitrary placement of 10 Au atoms on the ZnO(111) surface, the MO molecule was observed to lie flat on the surface. Similarly, Boumya et al. used MD simulations to study MO adsorption on the (110) surface of various metal chlorides and concluded that the dye adsorbs via van der Waals interactions, with a separation distance of over 3.0 Å [57].
To investigate MO adsorption on the Ag(111) surface, we performed MD simulations of an extended Ag–water interface featuring 1 anionic MO molecule ( MO ), 2 Na + , 1 BH 4 , and 2000 water molecules. After energy minimization, we performed an MD run of 8 ns at 300 K. Within the first 3 ns, the sulfonate group of MO migrated to the Ag(111) surface, and by 6 ns, the molecule adopted a flat orientation on the surface (Figure 6). Upon further prolongation of the MD simulation, no desorption of MO was observed. These findings support the commonly proposed mechanism in which Ag NPs facilitate electron transfer (from BH 4 ) by forming stable interactions with dye (MO) molecules.
To investigate the adsorption behavior of BH 4 ions on the Ag(111) surface, we introduced additional four BH 4 ions and four Na + counterions into the final configuration from the previous MD simulation. As shown in Figure 7a (snapshot at 19.45 ns), up to four BH 4 ions were adsorbed onto the Ag(111) surface and were observed to diffuse laterally along the surface, indicating favorable interactions between BH 4 and the Ag(111) surface.
To further characterize the adsorption of different species, radial distribution function (RDF) analysis was performed (Figure 7b,c). The RDF was normalized such that g(r) = 1 at r = 10 Å to allow direct comparison between different species. This distance was chosen because the number density of water was found to converge at around 10 Å (Figure S3), making it an appropriate reference point. The RDF of the nitrogen atom in the azo group of MO relative to the surface Ag atoms exhibited a sharp peak at 3.00 Å (Figure 7b, blue solid curve). In turn, Ag–O contacts with the oxygen atoms of nearby water molecules displayed a peak at a slightly greater distance of 3.13 Å (Figure 7b, red dashed curve).
For BH 4 ions, the RDF of hydride ( H ) atoms showed a pronounced peak at 3.27 Å from the Ag(111) surface (Figure 7b, green dashed curve), indicating association via van der Waals interactions. In contrast, Na + ions showed only occasional and transient contact to the Ag(111) surface, with only a weak RDF peak appearing at ~3.20 Å (Figure 7b, black dash-dotted curve). Indeed, Na + remains predominantly solvated in water rather than binding to the Ag surface.
We furthermore observed an Na + ion transiently approaches the sulfonate group of MO , as indicated by a pronounced RDF peak at a relatively large distance of ~4.70 Å (Figure 7c, black solid curve), arguably acting as a diffuse positive charge to neutralize the dye’s negative charge, rather than forming a direct, stable coordination. The residence time of Na + ion within 5 Å of the oxygen atoms of the sulfonate group was found to be only 5.4 ps (Figure S4), indicating a short-lived interaction. RDF analysis also confirmed negligible interaction between Na + and the azo group of MO (Figure 7c, cyan solid curve).
Similarly, BH 4 ions exhibited a low probability of approaching the azo group of MO at ~3.00 Å (Figure 7c, brown solid curve), indicating limited direct interaction. This likely reflects the Coulomb repulsion that prevents BH 4 from directly reacting with the azo group of MO , in contrast to the favorable electrostatic attraction between Na + and the sulfonate group. Overall, these MD simulation results support the general view of the MO degradation mechanism, in which Ag NPs simultaneously interact with both MO and BH 4 . In the absence of Ag NPs, these solute species would remain dispersed in solution, preventing effective electron transfer from BH 4 to MO . Thus, Ag NPs facilitate electron transfer without requiring direct contact between the reducing agent and the dye molecule.

2.6.2. Insights into the MO Degradation Mechanisms via DFT Calculations

While the MD simulations provided atomic-level insights into the adsorption behavior of MO dye on the Ag(111) surface, they could not capture the electronic details of the degradation reaction mechanism. To address the electron transfer between the two reactants, we therefore employed DFT calculations to investigate the possible reaction pathways underlying MO degradation.
As a first step, we aimed to identify the most reactive sites on the MO dye susceptible to the interaction with BH 4 . For this, condensed Fukui indices ( f + and f ) were calculated for each atom of the optimized MO geometry (Figure 8a,b). The Fukui function is a well-established theoretical tool for predicting site-specific chemical reactivity at the atomic level, where a high f + value indicates a site favorable for nucleophilic attack (electron donation), and a high f value corresponds to electrophilic susceptibility (electron acceptance) [58,59,60].
The analysis revealed that the N9 atom of the azo group possesses the highest f + value (Figure 8b), suggesting it as the primary site for nucleophilic hydride attack by BH 4 . This hydride addition is expected to initiate electron delocalization within the azo group, with the neighboring N8 atom serving as an electron acceptor. Additionally, the C6 atom on the aromatic ring emerged as a secondary possible site for nucleophilic attack, though with lower reactivity than N9. On the other hand, the N16 atom exhibited the highest f value, indicating its propensity to accept a proton ( H + ), potentially forming a protonated amine H–N CH 3 2 + moiety. However, this protonation does not directly contribute to the azo bond cleavage or degradation of MO. Overall, the condensed Fukui index analysis provides a valuable foundation for our subsequent DFT-based mechanistic study of MO degradation.
(A) First MO degradation. Based on the condensed Fukui index values, we investigated the initial step of MO degradation using DFT calculations within a continuum water solvent model. Taking the free energy of the MO anion as the reference, the association of a Na + ion with the sulfonyl group of MO slightly lowers the free energy by 0.6 kcal/mol. This suggests that Na + remains weakly bound and can easily dissociate, in agreement with the calculated residence time of 5.4 ps from the above MD simulations.
The subsequent hydride attack from BH 4 was explored via two possible pathways: (i) direct hydride transfer (TS1), defined by an imaginary frequency of −774.3 cm−1, and (ii) Na + -mediated hydride transfer (TS2), with an imaginary frequency of −673.0 cm−1, to the N9 atom of the azo group (Figure 9a). The direct hydride transfer proceeds with a lower activation energy (+37.4 kcal/mol) compared to the Na + -mediated pathway (+41.6 kcal/mol). Upon reaching the respective transition states (TS1 and TS2), the N–N bond length of the azo group increases from 1.27 Å to 1.33 Å (TS1) and 1.35 Å (TS2), indicating bond weakening (Figure 9b).
The lower activation energy for the direct hydride transfer (TS1) may be attributed to the higher aromaticity of the benzene ring attached to the SO 3 group, compared to TS2. Aromaticity in this study is qualitatively assessed by analyzing deviations in C–C bond lengths and the S1–C2–C5 angle (see atom numbering in Figure 8a) relative to the unperturbed MO structure. The average C–C bond length deviation in TS2 is 0.9%, slightly larger than that of TS1 (0.5%). Moreover, the S1–C2–C5 angle in TS2 deviates by 11.4% from MO , whereas in TS1, the deviation is only 0.9%. These structural distortions in TS2 are not sufficiently compensated by the greater charge transfer from BH 4 to N9 (−0.807e for TS2 vs. −0.711e for TS1, based on natural population charge analysis).
It is worth noting that when Na + was explicitly positioned at the sulfonyl group of MO and the direct hydride transfer transition state was sought, the optimization consistently converged to TS2, despite several attempts. After hydride transfer from BH 4 to the N9 atom, the resulting intermediate (IN3) exhibits a relative free energy of +25.9 kcal/mol. The N–N distance of the azo group in IN3 further elongates to 1.41 Å, indicating significant bond weakening, which makes the molecule susceptible to further degradation. These findings suggest that while Na + stabilizes charge in the system, its presence at a high concentration could inhibit MO degradation. This aligns with experimental observations showing that at higher Na BH 4 concentrations (~0.3 M), the apparent rate constant tends to decrease (see Table 2).
We also investigated the influence of a Ag atom by associating it with the azo group, as suggested by the MD simulations. This pathway exhibited an energy barrier of +38.3 kcal/mol, slightly higher than that of the direct hydride transfer (TS1) (Figure S5). These results imply that under the current model, Ag NPs primarily act as scaffolds to facilitate the proximity of MO and Na BH 4 , rather than directly participating in the electron transfer process during this step.
(B) Second MO degradation. The second step of MO degradation involves proton transfer processes that complete the breakdown of the MO molecule. To account for the effect of pH, the relative free energies (Figure 10a) were corrected to pH 3, 5, and 9 using the approach previously employed [61]. This allows a direct examination of how pH influences the energetics of the degradation pathway. Based on the optimal synthesis conditions for Ag NPs, the system pH (before the introduction of MO dye) is thus expected to be around 9. However, for the MO solution itself, the orangish hue suggests an acidic environment with a pH in the range of 3.1–4.4 [62]. Therefore, when the Ag NP dispersion and MO solution are mixed, the final pH is expected to be lower than 9.
Without pH correction (i.e., treating [ H + ] = 1 M, pH = 0; black line in Figure 10a), following the formation of the intermediate IN3, this negatively charged species readily accepts a proton from the solution to form IN4 (a hydrazine-like structure) with a slightly reduced free energy of −11.0 kcal/mol. Subsequent protonation yields IN5, which has a slightly higher free energy of −2.8 kcal/mol. The N–N bond length in IN5 elongates to 1.51 Å (Figure 10b), indicating further bond weakening and susceptibility to cleavage. The intermediate IN5 then interacts with BH 4 to undergo hydride transfer via TS3, with a low imaginary frequency of −267.1 cm−1, exhibiting a relative free-energy barrier of +24.8 kcal/mol, which is notably lower than that of the first step (+37.4 kcal/mol).
This hydride transfer yields the degradation products P1 (sodium 4-aminobenzenesulfonate) and P2 (N,N-dimethyl-p-phenylenediamine). These products correlate with the observed increase in the UV–Vis absorbance at 250 nm (Figure 3 and Figure 4) [63]. To confirm this, we conducted TD-DFT calculations to predict the UV–Vis spectra of MO, P1, and P2. As shown in Figure 11, the predicted spectra agree well with the experimental spectra in Figure 3 and Figure 4, both in the peak positions and the spectral shifts observed during MO degradation.
At slightly higher pH values of 3 and 5, the energy barriers from IN4 to TS3 increase to +28.8 and +31.6 kcal/mol, respectively, though they remain lower than the barrier observed in the first step (Figure 10a vs. Figure 9a). However, at a basic pH of 9 (Figure 10a, green line), the energy barrier increases significantly to approximately 37.0 kcal/mol, comparable to that of the first step (+37.4 kcal/mol). Therefore, at basic pH (>9), the second hydride transfer becomes the rate-determining step. This suggests that MO degradation efficiency decreases at a higher pH, which is consistent with the findings of Kgatle et al. [64], who observed a similar pH-dependent degradation trend using trimetallic Fe/Cu/Ag nanoparticles. This aspect warrants further systematic investigation.
In conclusion, the present DFT study provides detailed insights into the mechanism of MO degradation. At low to neutral pH, the first step—direct hydride transfer via TS1—remains the rate-determining step. However, at a higher pH (>9), the second step involving proton transfer is energetically less favorable and thus becomes the rate-determining step. These theoretical findings are consistent with experimental observations and highlight the critical influence of pH on the degradation efficiency of MO dye.

3. Materials and Methods

3.1. Chemicals and Reagents

All chemicals employed were analytical grade. Silver nitrate (Ag NO 3 ) purchased from Sigma-Aldrich (Darmstadt, Germany), sodium borohydride ( NaBH 4 ) from Thermo Scientific Chemicals (Waltham, MA, USA), and methyl orange powder ( C 14 H 14 N 3 NaO 3 S ) from LabChem (Auckland, New Zealand) were used as obtained. Sodium hydroxide (NaOH) from Ajax Finechem (New South Wales, Australia) was used as received. Deionized water with a resistance of 18 MΩ, taken from Scientific Promotion (Bangkok, Thailand) was used to prepare all aqueous solutions. All glasswares were cleaned with nitric acid, and then rinsed with deionized water prior to use.

3.2. Preparation of Nypa Fruticans Fruit Husk Extract and Green Synthesis of Ag NPs

The preparation of Nypa fruticans fruit husk (NF) extract and the synthesis of Ag NPs were carried out following our previous study [12]. Fresh NF husks were collected from the Pak Phanang River Basin, Nakhon Si Thammarat, Thailand. The husks were thoroughly washed with water, oven-dried at 60 °C for 42 h, and ground into a fine powder using an electric grinder (Philips Mixer Grinder HL1605). To obtain the extract, 10 g of NF powder was mixed with 100 mL of methanol and stirred continuously for 3 h. The mixture was then filtered through Whatman filter paper No. 1 (St. Louis, MO, USA), and the filtrate was concentrated using a vacuum evaporator at 50 °C. The obtained brown extract powder was stored at 4 °C for further applications.
For the green synthesis of Ag NPs, the method employed in our previous study was also used in the present study. In brief, a solution of 1.0 mL of 0.2% w/v NF extract was added dropwise to 20 mL of 1.0 mM silver nitrate under vigorous stirring at room temperature. The pH of the solution was adjusted to 9 using sodium hydroxide, and the reaction mixture was heated at 60 °C for 45 min. The formation of Ag NPs was confirmed by a color change to yellow, indicating the presence of reduced silver ions. It is noteworthy that the Ag NPs synthesized in this work were freshly prepared using the same synthesis method as in our previous study [12].

3.3. Characterization of Ag NPs

The Ag NP colloidal solution was analyzed using a SPECTROstar Nano spectrophotometer (BMG LABTECH, Ortenberg, Germany) within the wavelength range of 300–800 nm to determine peak absorbance and characteristic surface plasmon resonance (SPR). The morphology and size of the nanoparticles were examined using Transmission Electron Microscopy (TEM; Talos F200i instrument, Thermo Fisher Scientific, Brno, Czech Republic) and Scanning Electron Microscopy (SEM; Apreo, FEI, Thermo Fisher Scientific, Brno, Czech Republic). Energy-Dispersive X-ray Spectroscopy (EDX; X-Max, Oxford. Instrument, UK) was utilized to confirm the formation of Ag NPs. Additionally, Fourier Transform Infrared Spectroscopy (FT-IR; INVENIO-S, Bruker, Germany) was performed to identify the functional groups present in both the NF extract and the synthesized Ag NPs.

3.4. Catalytic Tests for Methyl Orange Dye Degradation

The catalytic activity of the synthesized Ag NPs was evaluated using the reduction of methyl orange, a model azo dye, in the presence of NaBH 4 . In each test, 2.5 mL of 0.1 mM MO solution was mixed with 0.5 mL of 0.2 M NaBH 4 , followed by the addition of varying volumes (25–100 μL) of the Ag NP solution. The MO solution was prepared using deionized water. For the commercial synthetic dye, solutions were prepared using either deionized water or real water samples, non-potable tap water collected at Nakhon Si Thammarat Rajabhat University (two independent samples from different locations; no pH adjustment) prior to the degradation tests. The degradation process was monitored by recording the UV–Vis absorption spectra of the reaction mixture in the wavelength range of 200–700 nm at regular time intervals, until the characteristic orange color of MO completely faded, indicated by the disappearance of the peak at 465 nm. The obtained data were used to analyze the degradation kinetics. Control experiments under identical conditions but without the addition of Ag NPs were also conducted to confirm that the observed dye degradation was attributed to the catalytic activity of the Ag NPs.

3.5. Quantum Chemical Calculations

All quantum chemical calculations were performed using the Gaussian 09 package [65]. Geometry optimizations for methyl orange (MO) and other relevant intermediates and transition states were conducted using the B3LYP functional [66,67] with the 6-31++G(d,p) basis set. In what follows, the MO molecule without the Na + counterion is denoted as MO . Electronic self-consistency used the following convergence criteria: root mean square (RMS) density matrix change of 1.0 × 10−8, maximum density matrix change of 1.0 × 10−6, and energy change of 1.0 × 10−6 Hartree. Geometry optimizations were iterated until the default force and displacement thresholds were satisfied as follows: maximum force of 4.5 × 10−4 Hartree/Bohr, RMS force of 3.0 × 10−4 Hartree/Bohr, maximum displacement of 1.8 × 10−3 Bohr, and RMS displacement of 1.2 × 10−3 Bohr. Frequency calculations were performed to obtain free-energy corrections and to confirm each stationary point as either an intermediate (no imaginary frequencies) or a transition state (a single imaginary frequency) [68].
Subsequently, single-point calculations were conducted at a higher level of theory, B3LYP-D3/6-311++G(d,p), using the polarizable continuum model (PCM) to mimic solvation in water. The D3 correction by Grimme et al. [69] was included to account for electronic dispersion interactions. Partial atomic charges were derived from the natural population analysis (NPA) [70]. Condensed Fukui functions were computed based on Hirshfeld charges, which have been demonstrated as effective tools for identifying reactive sites [59,60,71,72]. Moreover, time-dependent DFT (TD-DFT) calculations were performed to predict the UV–Vis spectra, employing the same functional, basis set, and PCM solvation model as in the single-point calculations.
The optimized gas-phase structures were subjected to restrained electrostatic potential (RESP) fitting to determine partial atomic charges. To maintain consistency with the standard General Amber Force Field (GAFF) approach [73] used in the molecular dynamics simulations, assessment of the partial charges is based on single-point calculations at the HF/6-31+G(d) level.

3.6. Molecular Dynamics Simulations

Molecular dynamics (MD) simulations were conducted with the Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS) software package (23 Jun 2022—Update 4) [74], utilizing a 1 fs time step for numerical integration. For bulk phase simulations, shifted-force potentials with a 12 Å cutoff were applied. To approximate Ewald summation for the long-range Coulomb interactions, the damped-shifted force potential with a damping parameter of 0.05 Å−1 was employed [75]. The simulations were carried out in the canonical (NVT) ensemble with a Nosé–Hoover thermostat to control the temperature.
A slab model of the Ag(111) surface, as used in our previous study [12], was constructed using the VESTA program [76]. The unit cell was replicated 6 × 11 × 3 times, resulting in a slab size of 30.0 × 31.8 × 21.2 Å3. During the MD simulations, we fixed the three bottom layers and applied 2D periodic boundary conditions to the Ag slab, with free boundary conditions in the [111] direction.
For modeling the Ag–water interface, a film of 2000 H 2 O molecules placed on the Ag slab was employed in analogy to a previous study of heavy metal association with Ag(111) [12]. Prior to the production runs, the system containing 2 Na + , 1 BH 4 , and 1 MO in 2000 H 2 O was carefully relaxed to accommodate the interface with the Ag(111) slab. Initial energy minimization was performed using the conjugate gradient (CG) algorithm [77]. This was followed by a 4 ns run at 300 K in the NVT ensemble. Next, additional 4 Na + and 4 BH 4 ions were randomly introduced into the final configuration, ensuring a minimum distance of 0.5 Å from all existing atoms. Another round of CG minimization was performed, followed by a 24 ns simulation at 300 K, with structural properties sampled from the last 4 ns.
To further illustrate the validity of the water model, we computed the number density profile of H 2 O along the surface normal (Figure S3). In the bulk-like region of the slab, the density reached a stable plateau of ~0.10 atoms/Å3, equivalent to ~0.033 molecules/Å3, which converts to ~0.996 g/cm3. This value is in excellent agreement with the experimental bulk water density (0.9965 g/cm3 at 298 K). As expected, oscillations were observed near the Ag surface due to interfacial layering, while a gradual density decay occurred at the liquid–vacuum interface [78,79].
The Ag-Ag interaction potential for the silver metal was taken from the Embedded Atom Model (EAM) [80]. In turn, interactions with C, O, and H atoms were described using GAFF2 [73] force fields. For this, the Lennard-Jones parameters for Ag were taken from the study by Heinz et al. [81], while those for Na + were from Mamatkulov et al.’s study [82]. For mixing the van der Waals parameters, the Lorentz–Berthelot combination rules were applied.

4. Conclusions

This study demonstrates the application of Ag NPs green-synthesized using Nypa fruticans fruit husk extract as a reducing and stabilizing agent. These Ag NPs were well-dispersed with an average diameter of approximately 4 nm, as revealed by TEM. The presence of biomolecules from the extract, notably hydroxyl (–OH) groups, on Ag NPs was confirmed by FT-IR spectroscopy. Under the optimal synthesis condition (pH 9), surface groups are predominantly deprotonated, imparting a negative surface charge on the Ag NPs, as previously reported [12], which prevents uncontrolled aggregation.
Prior to their application in dye degradation, reaction parameters were systematically optimized. Our synthesized Ag NPs efficiently catalyzed the degradation of methyl orange dye in the presence of Na BH 4 , achieving complete degradation within one minute. UV–Vis spectral analysis showed that the degradation process followed pseudo-first-order kinetics, with high rate constants indicating efficient catalytic activity.
The mechanism of MO degradation was elucidated through a combined approach of MD and DFT calculations, providing novel insights into the catalytic process. MD simulations suggested preferential adsorption of MO via its azo group onto the Ag(111) surface and the proximity of Na + ions near the sulfonyl groups. DFT calculations, incorporating condensed Fukui function analysis, revealed that the initial step involves direct hydride transfer from BH 4 to the N atom of the azo group via a low-energy transition state (TS1), which is more favorable than the Na + -mediated pathway (TS2). This first step represents the rate-determining process under neutral and acidic conditions. The subsequent step of MO degradation was also evaluated theoretically, revealing pH-dependent energy profiles due to proton involvement. Under alkaline conditions (at basic pH, >9), the second hydride transfer step becomes the rate-determining step, potentially reducing the degradation efficiency. These findings are consistent with previous experimental reports on the pH-dependent kinetics of MO degradation and provide additional mechanistic insight into the catalytic process. Further systematic experimental investigations are recommended to confirm the simulation-predicted effects of pH on the catalytic performance of Ag NPs.
Overall, this study highlights the potential of Nypa fruticans fruit husk extract for the sustainable synthesis of Ag NPs with excellent catalytic performance in dye degradation. Moreover, the theoretical insights gained from MD and DFT analyses complement the experimental results, offering a comprehensive understanding of the degradation mechanism and the influence of pH on the process. These insights could guide the design of more efficient catalytic systems for environmental remediation applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30183738/s1, Figure S1: comparison of the UV–Vis spectra of the Ag NPs synthesized in this study (orange curve) and in our previous report (Wonglakhon et al., 2025 [12], blue curve). Both batches exhibit a characteristic SPR band at 424 nm. The peak absorbance at 424 nm is 1.6415 (this study) vs. 1.6493 (previous study), indicating < 0.5% difference; Figure S2: UV–Vis spectral changes for the reduction of the commercial synthetic dye by the synthesized Ag NPs in real water. Spectra were recorded in non-potable tap water from different locations: left, real water sample I; right, real water sample II. Legends indicate degradation time and the corresponding % degradation. For both samples, degradation reached ~83% at 5 min and ~86% by 6 min, accompanied by decolorization from deep yellow to nearly colorless, indicating a modest matrix effect while maintaining high catalytic activity; Figure S3: number density profile of water (in Å−3) along surface normal of the Ag(111) surface. A red dashed line marks the position where the number density reaches a plateau, indicating convergence. This distance was used as a reference point for normalizing the radial distribution functions; Figure S4: survival probability of a sodium ion within 5 Å of the oxygen atoms of the sulfonate group of MO dye. An exponential decay function was fitted to the data, providing an average residence time of 5.39 ps; Figure S5: relative free-energy profile (in kcal/mol) for the first step of MO degradation via (i) a direct hydride transfer (TS1), (ii) Na+-mediated hydride transfer (TS2), and (iii) direct hydride transfer with the incorporation of Ag atom (TS1-Ag)—from BH 4 to N9 atom of the azo group. The 3D-optimized structures MO-Ag and TS1-Ag, with distance labels in Å.

Author Contributions

T.W.: Conceptualization, investigation, theoretical study, visualization, methodology, formal analysis, supervision, project administration, writing—original draft, writing—review and editing, and funding acquisition. A.C.: Investigation. P.N.: Writing—review and editing. B.N.: Writing—review and editing. D.Z.: Supervision, writing—review and editing. Y.T.: Conceptualization, investigation, visualization, methodology, formal analysis, supervision, project administration, validation, writing—original draft, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Walailak University under the New Researcher Development scheme (Contract Number WU68262).

Data Availability Statement

Data are contained within the article or Supplementary Materials. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

All the authors declare that they have no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
Ag NPsSilver nanoparticles
SPRSurface plasmon resonance
NFNypa fruticans fruit husk
Ag NO 3 Silver nitrate
Na BH 4 Sodium borohydride
NaOHSodium hydroxide
MOMethyl orange
TEMTransmission Electron Microscopy
SEMScanning Electron Microscopy
EDXEnergy-Dispersive X-ray Spectroscopy
FT-IRFourier Transform Infrared Spectroscopy
DFTDensity functional theory
TD-DFTTime-dependent density functional theory
PCMPolarizable continuum model
NPANatural population analysis
RMSRoot mean square
MDMolecular dynamics
RDFRadial distribution function
CGConjugate gradient
RESPRestrained electrostatic potential
GAFFGeneral Amber Force Field
EAMEmbedded Atom Model

References

  1. Kishor, R.; Purchase, D.; Saratale, G.D.; Saratale, R.G.; Ferreira, L.F.R.; Bilal, M.; Chandra, R.; Bharagava, R.N. Ecotoxicological and Health Concerns of Persistent Coloring Pollutants of Textile Industry Wastewater and Treatment Approaches for Environmental Safety. J. Environ. Chem. Eng. 2021, 9, 105012. [Google Scholar] [CrossRef]
  2. Carney Almroth, B.; Cartine, J.; Jönander, C.; Karlsson, M.; Langlois, J.; Lindström, M.; Lundin, J.; Melander, N.; Pesqueda, A.; Rahmqvist, I.; et al. Assessing the Effects of Textile Leachates in Fish Using Multiple Testing Methods: From Gene Expression to Behavior. Ecotoxicol. Environ. Saf. 2021, 207, 111523. [Google Scholar] [CrossRef] [PubMed]
  3. Tkaczyk, A.; Mitrowska, K.; Posyniak, A. Synthetic Organic Dyes as Contaminants of the Aquatic Environment and Their Implications for Ecosystems: A Review. Sci. Total Environ. 2020, 717, 137222. [Google Scholar] [CrossRef]
  4. Szyguła, A.; Guibal, E.; Palacín, M.A.; Ruiz, M.; Sastre, A.M. Removal of an Anionic Dye (Acid Blue 92) by Coagulation–Flocculation Using Chitosan. J. Environ. Manag. 2009, 90, 2979–2986. [Google Scholar] [CrossRef]
  5. Maxwell, J.C.; Baker, B.C.; Faul, C.F.J. Controlled Removal of Organic Dyes from Aqueous Systems Using Porous Cross-Linked Conjugated Polyanilines. ACS Appl. Polym. Mater. 2023, 5, 662–671. [Google Scholar] [CrossRef] [PubMed]
  6. Samari, F.; Salehipoor, H.; Eftekhar, E.; Yousefinejad, S. Low-Temperature Biosynthesis of Silver Nanoparticles Using Mango Leaf Extract: Catalytic Effect, Antioxidant Properties, Anticancer Activity and Application for Colorimetric Sensing. New J. Chem. 2018, 42, 15905–15916. [Google Scholar] [CrossRef]
  7. Ertürk, A.S. Biosynthesis of Silver Nanoparticles Using Epilobium Parviflorum Green Tea Extract: Analytical Applications to Colorimetric Detection of Hg2+ Ions and Reduction of Hazardous Organic Dyes. J. Clust. Sci. 2019, 30, 1363–1373. [Google Scholar] [CrossRef]
  8. Kadam, J.; Dhawal, P.; Barve, S.; Kakodkar, S. Green Synthesis of Silver Nanoparticles Using Cauliflower Waste and Their Multifaceted Applications in Photocatalytic Degradation of Methylene Blue Dye and Hg2+ Biosensing. SN Appl. Sci. 2020, 2, 738. [Google Scholar] [CrossRef]
  9. Islam, M.A.; Jacob, M.V.; Antunes, E. A Critical Review on Silver Nanoparticles: From Synthesis and Applications to Its Mitigation through Low-Cost Adsorption by Biochar. J. Environ. Manag. 2021, 281, 111918. [Google Scholar] [CrossRef]
  10. Moond, M.; Singh, S.; Sangwan, S.; Devi, P.; Beniwal, A.; Rani, J.; Kumari, A.; Rani, S. Biosynthesis of Silver Nanoparticles Utilizing Leaf Extract of Trigonella Foenum-Graecum L. for Catalytic Dyes Degradation and Colorimetric Sensing of Fe3+/Hg2+. Molecules 2023, 28, 951. [Google Scholar] [CrossRef]
  11. Magdy, G.; Aboelkassim, E.; Abd Elhaleem, S.M.; Belal, F. A Comprehensive Review on Silver Nanoparticles: Synthesis Approaches, Characterization Techniques, and Recent Pharmaceutical, Environmental, and Antimicrobial Applications. Microchem. J. 2024, 196, 109615. [Google Scholar] [CrossRef]
  12. Wonglakhon, T.; Jommala, N.; Laksee, S.; Nuengmatcha, P.; Ninwong, B.; Zahn, D.; Thepchuay, Y. Experimental and Computational Study of Ecofriendly Synthesis of Silver Nanoparticles from Natural Extracts: Self-Controlled Nucleation and Growth, and Colorimetric Detection of Heavy Metal Ions. Surf. Interfaces 2025, 68, 106618. [Google Scholar] [CrossRef]
  13. Akpakpan, A.E.; Akpabio, U.D.; Obot, I.B. Evaluation of Physicochemical Properties and Soda Pulping of Nypa Fruticans Frond and Petiole. Elixir Appl. Chem. 2012, 45, 7664–7668. [Google Scholar]
  14. Nypa Fruticans (Nipa Palm). CABI Compendium. 2022. Available online: https://www.cabidigitallibrary.org/doi/abs/10.1079/cabicompendium.36772 (accessed on 11 August 2025).
  15. Cheablam, O.; Chanklap, B. Sustainable Nipa Palm (Nypa Fruticans Wurmb.) Product Utilization in Thailand. Scientifica 2020, 2020, 3856203. [Google Scholar] [CrossRef]
  16. Axon, S.; James, D. The UN Sustainable Development Goals: How Can Sustainable Chemistry Contribute? A View from the Chemical Industry. Curr. Opin. Green Sustain. Chem. 2018, 13, 140–145. [Google Scholar] [CrossRef]
  17. Lee, S.H.; Jun, B.H. Silver Nanoparticles: Synthesis and Application for Nanomedicine. Int. J. Mol. Sci. 2019, 20, 865. [Google Scholar] [CrossRef] [PubMed]
  18. Li, M.; Li, J.; Qin, C.; Guo, X.; Wang, H.; Zeng, Z.; Yuan, X. Cuprous-Mediated Peroxymonosulfate Activation for Fenton-like Removal of Micropollutants: The Function of Co-Catalyst and the Accelerated Degradation Mechanism. Ecotoxicol. Environ. Saf. 2023, 264, 115435. [Google Scholar] [CrossRef]
  19. Zhu, K.; Zhang, R.; Yao, Y.; Zhao, M.; Huang, X.; Huang, Z.; Yang, Y.; Liang, X.; Yan, K. Oxygen Vacancies Engineering in Iron-Doped Nickel Molybdate Nanorods to Accelerate Electron Transfer for Rapid Organic Pollutants Removal. Sep. Purif. Technol. 2025, 372, 133391. [Google Scholar] [CrossRef]
  20. Vidhu, V.K.; Philip, D. Catalytic Degradation of Organic Dyes Using Biosynthesized Silver Nanoparticles. Micron 2014, 56, 54–62. [Google Scholar] [CrossRef]
  21. Joseph, S.; Mathew, B. Microwave-Assisted Green Synthesis of Silver Nanoparticles and the Study on Catalytic Activity in the Degradation of Dyes. J. Mol. Liq. 2015, 204, 184–191. [Google Scholar] [CrossRef]
  22. Meenakumari, M.; Philip, D. Degradation of Environment Pollutant Dyes Using Phytosynthesized Metal Nanocatalysts. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2015, 135, 632–638. [Google Scholar] [CrossRef]
  23. Edison, T.N.J.I.; Atchudan, R.; Sethuraman, M.G.; Lee, Y.R. Reductive-Degradation of Carcinogenic Azo Dyes Using Anacardium Occidentale Testa Derived Silver Nanoparticles. J. Photochem. Photobiol. B 2016, 162, 604–610. [Google Scholar] [CrossRef]
  24. Bogireddy, N.K.R.; Kiran Kumar, H.A.; Mandal, B.K. Biofabricated Silver Nanoparticles as Green Catalyst in the Degradation of Different Textile Dyes. J. Environ. Chem. Eng. 2016, 4, 56–64. [Google Scholar] [CrossRef]
  25. Varadavenkatesan, T.; Selvaraj, R.; Vinayagam, R. Phyto-Synthesis of Silver Nanoparticles from Mussaenda Erythrophylla Leaf Extract and Their Application in Catalytic Degradation of Methyl Orange Dye. J. Mol. Liq. 2016, 221, 1063–1070. [Google Scholar] [CrossRef]
  26. Jyoti, K.; Singh, A. Green Synthesis of Nanostructured Silver Particles and Their Catalytic Application in Dye Degradation. J. Genet. Eng. Biotechnol. 2016, 14, 311–317. [Google Scholar] [CrossRef]
  27. Sengan, M.; Veeramuthu, D.; Veerappan, A. Photosynthesis of Silver Nanoparticles Using Durio Zibethinus Aqueous Extract and Its Application in Catalytic Reduction of Nitroaromatics, Degradation of Hazardous Dyes and Selective Colorimetric Sensing of Mercury Ions. Mater. Res. Bull. 2018, 100, 386–393. [Google Scholar] [CrossRef]
  28. Ismail, M.; Gul, S.; Khan, M.I.; Khan, M.A.; Asiri, A.M.; Khan, S.B. Medicago Polymorpha-Mediated Antibacterial Silver Nanoparticles in the Reduction of Methyl Orange. Green Process. Synth. 2019, 8, 118–127. [Google Scholar] [CrossRef]
  29. Santhosh, A.S.; Sandeep, S.; Kumara Swamy, N. Green Synthesis of Nano Silver from Euphorbia Geniculata Leaf Extract: Investigations on Catalytic Degradation of Methyl Orange Dye and Optical Sensing of Hg2+. Surf. Interfaces 2019, 14, 50–54. [Google Scholar] [CrossRef]
  30. Raina, S.; Roy, A.; Bharadvaja, N. Degradation of Dyes Using Biologically Synthesized Silver and Copper Nanoparticles. Environ. Nanotechnol. Monit. Manag. 2020, 13, 100278. [Google Scholar] [CrossRef]
  31. Edison, T.N.J.I.; Atchudan, R.; Karthik, N.; Balaji, J.; Xiong, D.; Lee, Y.R. Catalytic Degradation of Organic Dyes Using Green Synthesized N-Doped Carbon Supported Silver Nanoparticles. Fuel 2020, 280, 118682. [Google Scholar] [CrossRef]
  32. Raj, S.; Singh, H.; Trivedi, R.; Soni, V. Biogenic Synthesis of AgNPs Employing Terminalia Arjuna Leaf Extract and Its Efficacy towards Catalytic Degradation of Organic Dyes. Sci. Rep. 2020, 10, 9616. [Google Scholar] [CrossRef] [PubMed]
  33. Chandra Paul, S.; Bhowmik, S.; Rani Nath, M.; Islam, M.S.; Kanti Paul, S.; Neazi, J.; Sabnam Binta Monir, T.; Dewanjee, S.; Abdus Salam, M. Silver Nanoparticles Synthesis in a Green Approach: Size Dependent Catalytic Degradation of Cationic and Anionic Dyes. Orient. J. Chem. 2020, 36, 353–360. [Google Scholar] [CrossRef]
  34. Shah, Z.; Gul, T.; Ali Khan, S.; Shaheen, K.; Anwar, Y.; Suo, H.; Ismail, M.; Alghamdi, K.M.; Salman, S.M. Synthesis of High Surface Area AgNPs from Dodonaea Viscosa Plant for the Removal of Pathogenic Microbes and Persistent Organic Pollutants. Mater. Sci. Eng. B 2021, 263, 114770. [Google Scholar] [CrossRef]
  35. Sarkar, M.; Denrah, S.; Das, M.; Das, M. Statistical Optimization of Bio-Mediated Silver Nanoparticles Synthesis for Use in Catalytic Degradation of Some Azo Dyes. Chem. Phys. Impact 2021, 3, 100053. [Google Scholar] [CrossRef]
  36. Kim, B.; Song, W.C.; Park, S.Y.; Park, G. Green Synthesis of Silver and Gold Nanoparticles via Sargassum Serratifolium Extract for Catalytic Reduction of Organic Dyes. Catalysts 2021, 11, 347. [Google Scholar] [CrossRef]
  37. Chikkanayakanahalli Paramesh, C.; Halligudra, G.; Gangaraju, V.; Sriramoju, J.B.; Shastri, M.; Kachigere, B.H.; Habbanakuppe, D.P.; Rangappa, D.; Kanchugarakoppal Subbegowda, R.; Doddakunche Shivaramu, P. Silver Nanoparticles Synthesized Using Saponin Extract of Simarouba Glauca Oil Seed Meal as Effective, Recoverable and Reusable Catalyst for Reduction of Organic Dyes. Results Surf. Interfaces 2021, 3, 100005. [Google Scholar] [CrossRef]
  38. Ibrahim, S.; Ahmad, Z.; Manzoor, M.Z.; Mujahid, M.; Faheem, Z.; Adnan, A. Optimization for Biogenic Microbial Synthesis of Silver Nanoparticles through Response Surface Methodology, Characterization, Their Antimicrobial, Antioxidant, and Catalytic Potential. Sci. Rep. 2021, 11, 770. [Google Scholar] [CrossRef]
  39. Rajasekar, R.; Thanasamy, R.; Samuel, M.; Edison, T.N.J.I.; Raman, N. Ecofriendly Synthesis of Silver Nanoparticles Using Heterotheca Subaxillaris Flower and Its Catalytic Performance on Reduction of Methyl Orange. Biochem. Eng. J. 2022, 187, 108447. [Google Scholar] [CrossRef]
  40. Vankdoth, S.; Velidandi, A.; Sarvepalli, M.; Vangalapati, M. Role of Plant (Tulasi, Neem and Turmeric) Extracts in Defining the Morphological, Toxicity and Catalytic Properties of Silver Nanoparticles. Inorg. Chem. Commun. 2022, 140, 109476. [Google Scholar] [CrossRef]
  41. Khan, W.; Khan, N.; Jamila, N.; Masood, R.; Minhaz, A.; Amin, F.; Atlas, A.; Nishan, U. Antioxidant, Antibacterial, and Catalytic Performance of Biosynthesized Silver Nanoparticles of Rhus Javanica, Rumex Hastatus, and Callistemon Viminalis. Saudi J. Biol. Sci. 2022, 29, 894–904. [Google Scholar] [CrossRef]
  42. Song, W.C.; Kim, B.; Park, S.Y.; Park, G.; Oh, J.W. Biosynthesis of Silver and Gold Nanoparticles Using Sargassum Horneri Extract as Catalyst for Industrial Dye Degradation. Arab. J. Chem. 2022, 15, 104056. [Google Scholar] [CrossRef]
  43. Sarkar, P.; Paul, S.K.; Sahu, K.; Mandal, S.; Islam, Q.A.; Chakrabarty, R. A Greener Synthesis of Silver Nanoparticles Using Tea Leaf and Aloe Vera Leaf Extract for Catalytic Organic Dye Degradation Application. ChemistrySelect 2023, 8, e202301906. [Google Scholar] [CrossRef]
  44. Desai, R.; Yadav, V.K.; Patel, B.; Rami, E.; Almoallim, H.S.; Ansari, M.J.; Choudhary, N.; Sahoo, D.K.; Patel, A. Assessment of Antimicrobial Activity and Methyl Orange Dye Removal by Klebsiella Pneumoniae-Mediated Silver Nanoparticles. Green Process. Synth. 2024, 13, 20240084. [Google Scholar] [CrossRef]
  45. Biswas, K.; Ahamed, Z.; Dutta, T.; Mallick, B.; Khuda-Bukhsh, A.R.; Biswas, J.K.; Mandal, S.K. Green Synthesis of Silver Nanoparticles from Waste Leaves of Tea (Camellia Sinensis) and Their Catalytic Potential for Degradation of Azo Dyes. J. Mol. Struct. 2024, 1318, 139448. [Google Scholar] [CrossRef]
  46. Yang, Z.; Gong, X.; Hu, Y.; Yue, P.; Lü, B.; Peng, F. Green Synthesis of Lichenan-Decorated Silver Nanoparticles for Catalytic Hydrogenation of Organic Dyes and Bacterial Disinfection. Chem. Eng. J. 2024, 487, 150516. [Google Scholar] [CrossRef]
  47. Moond, M.; Singh, S.; Rani, J.; Beniwal, A.; Sharma, R.K. Bio-Fabricated Silver Nanoparticles for Catalytic Degradation of Toxic Dyes and Colorimetric Sensing of Hg2+. ChemistrySelect 2024, 9, e202401826. [Google Scholar] [CrossRef]
  48. Hijazi, B.U.; Faraj, M.; Mhanna, R.; El-Dakdouki, M.H. Biosynthesis of Silver Nanoparticles as a Reliable Alternative for the Catalytic Degradation of Organic Dyes and Antibacterial Applications. Curr. Res. Green Sustain. Chem. 2024, 8, 100408. [Google Scholar] [CrossRef]
  49. Kurra, H.; Velidandi, A.; Pabbathi, N.P.P.; Godishala, V. Aqueous Cymbopogon Citratus Extract Mediated Silver Nanoparticles: Part II. Dye Degradation Studies. Eng 2025, 6, 102. [Google Scholar] [CrossRef]
  50. Kaliraja, T.; Kalla, R.M.N.; Al-Zahrani, F.A.M.; Vattikuti, S.V.P.; Lee, J. Eco-Friendly Synthesis of Silver Nanoparticles from Ligustrum Ovalifolium Flower and Their Catalytic Applications. Nanomaterials 2025, 15, 1087. [Google Scholar] [CrossRef]
  51. Priyadarshini, V.; Tharini, K.; Kalaimagal, G.; Alvin Kalicharan, A.; Subhashini, B.; Rathinavelu, A.; Mohan, S. Green Engineered Silver (Ag NPs) Nanoparticles Enable for Thermal, Optical Behavior and Catalytic Elimination of Organic Pollutants. Results Surf. Interfaces 2025, 20, 100593. [Google Scholar] [CrossRef]
  52. Goswami, P.J.; Kashyap, K.; Dutta, D.; Lal, M.; Hmar, J.J.L.; Bori, J. A Facile Green Synthesis of Ultra Small Silver Nanoparticles Using Aqueous Branch Extract of Dipterocarpus Retusus (Hollong) Promoting Catalytic Degradation of Organic Pollutants for Environmental Remediation. Ind. Crops Prod. 2025, 235, 121689. [Google Scholar] [CrossRef]
  53. Basappa, K.S.; Raghava, S.; Umesha, S. Biogenic Silver Nanoparticles from Annona Reticulata Endophyte: Antibacterial, Anticancer, and Dye Degradation Potential. Vegetos 2025. [Google Scholar] [CrossRef]
  54. Lima, A.K.O.; Vieira, Í.R.S.; Souza, L.M.d.S.; Florêncio, I.; Silva, I.G.M.d.; Tavares Junior, A.G.; Machado, Y.A.A.; Santos, L.C.d.; Taube, P.S.; Nakazato, G.; et al. Green Synthesis of Silver Nanoparticles Using Paullinia Cupana Kunth Leaf Extract Collected in Different Seasons: Biological Studies and Catalytic Properties. Pharmaceutics 2025, 17, 356. [Google Scholar] [CrossRef] [PubMed]
  55. Chetry, L.; Sarma, P.P.; Baruah, P.K. Zingiber Sianginensis-Mediated Green Synthesis of Multifunctional Silver Nanoparticles for Antimicrobial, Antioxidant, Heavy Metal Sensing, and Environmental Remediation. Colloids Surf. A Physicochem. Eng. Asp. 2025, 720, 137156. [Google Scholar] [CrossRef]
  56. Fahimirad, B.; Malekshah, R.E.; Chamjangali, M.A.; Abasabadi, R.K.; Bromand, S. Theoretical and Experimental Study of the Photodegradation of Methyl Orange in the Presence of Different Morphologies of Au-ZnO Using Monte Carlo Dynamic Simulation. Environ. Sci. Pollut. Res. 2022, 29, 55131–55146. [Google Scholar] [CrossRef]
  57. Boumya, W.; Khnifira, M.; Farid, Z.; Sadiq, M.; Elhalil, A.; Achak, M.; Kaya, S.; Barka, N.; Abdennouri, M. Comparative Study of Cationic Nile Blue and Anionic Methyl Orange Dyes Adsorption in Water on the (110) Surface of Metal Chlorides by DFT and MD Approaches. J. Phys. Chem. Solids 2024, 185, 111738. [Google Scholar] [CrossRef]
  58. Qutob, M.; Doğan, Ş.; Rafatullah, M. Heterogeneous Activation of Persulfate by Activated Carbon for Efficient Acetaminophen Degradation: Mechanism, Kinetics, Mineralization, and Density Functional Theory. ChemistrySelect 2022, 7, e202201249. [Google Scholar] [CrossRef]
  59. Oláh, J.; Alsenoy, C.V.; Sannigrahi, A.B. Condensed Fukui Functions Derived from Stockholder Charges: Assessment of Their Performance as Local Reactivity Descriptors. J. Phys. Chem. A 2002, 106, 3885–3890. [Google Scholar] [CrossRef]
  60. Roy, R.K. Stockholders Charge Partitioning Technique. A Reliable Electron Population Analysis Scheme to Predict Intramolecular Reactivity Sequence. J. Phys. Chem. A 2003, 107, 10428–10434. [Google Scholar] [CrossRef]
  61. Wonglakhon, T.; Surawatanawong, P. Mechanistic Insights into HCO2H Dehydrogenation and CO2 Hydrogenation Catalyzed by Ir(Cp*) Containing Tetrahydroxy Bipyrimidine Ligand: The Role of Sodium and Proton Shuttle. Dalton Trans. 2018, 47, 17020–17031. [Google Scholar] [CrossRef]
  62. Bolton, P.D.; Ellis, J.; Fleming, K.A.; Lantzke, I.R. Protonation of Azobenzene Derivatives. I. Methyl Orange and Ortho-Methyl Orange. Aust. J. Chem. 1973, 26, 1005–1014. [Google Scholar] [CrossRef]
  63. Kim, H.J.; Jung, Y.J.; Son, S.H.; Choi, W.S. Compressible Separator and Catalyst for Simultaneous Separation and Purification of Emulsions and Aqueous Pollutants. ACS Omega 2023, 8, 40741–40753. [Google Scholar] [CrossRef]
  64. Kgatle, M.; Sikhwivhilu, K.; Ndlovu, G.; Moloto, N. Degradation Kinetics of Methyl Orange Dye in Water Using Trimetallic Fe/Cu/Ag Nanoparticles. Catalysts 2021, 11, 428. [Google Scholar] [CrossRef]
  65. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  66. Becke, A.D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  67. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed]
  68. Fukui, K. The Path of Chemical Reactions—The IRC Approach. Acc. Chem. Res. 1981, 14, 363–368. [Google Scholar] [CrossRef]
  69. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  70. Reed, A.E.; Weinstock, R.B.; Weinhold, F. Natural Population Analysis. J. Chem. Phys. 1985, 83, 735–746. [Google Scholar] [CrossRef]
  71. Wang, B.; Rong, C.; Chattaraj, P.K.; Liu, S. A Comparative Study to Predict Regioselectivity, Electrophilicity and Nucleophilicity with Fukui Function and Hirshfeld Charge. Theor. Chem. Acc. 2019, 138, 124. [Google Scholar] [CrossRef]
  72. Liu, S.; Rong, C.; Lu, T. Information Conservation Principle Determines Electrophilicity, Nucleophilicity, and Regioselectivity. J. Phys. Chem. A 2014, 118, 3698–3704. [Google Scholar] [CrossRef]
  73. Wang, J.; Wolf, R.M.; Caldwell, J.W.; Kollman, P.A.; Case, D.A. Development and Testing of a General Amber Force Field. J. Comput. Chem. 2004, 25, 1157–1174. [Google Scholar] [CrossRef]
  74. Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular Dynamics. J. Comput. Phys. 1995, 117, 1–19. [Google Scholar] [CrossRef]
  75. Zahn, D.; Schilling, B.; Kast, S.M. Enhancement of the Wolf Damped Coulomb Potential: Static, Dynamic, and Dielectric Properties of Liquid Water from Molecular Simulation. J. Phys. Chem. B 2002, 106, 10725–10732. [Google Scholar] [CrossRef]
  76. Momma, K.; Izumi, F. VESTA 3 for Three-Dimensional Visualization of Crystal, Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  77. Polyak, B.T. The Conjugate Gradient Method in Extremal Problems. USSR Comput. Math. Math. Phys. 1969, 9, 94–112. [Google Scholar] [CrossRef]
  78. Jin, J.; Wang, X.; Wick, C.D.; Dang, L.X.; Miller, J.D. Silica Surface States and Their Wetting Characteristics. Surf. Innov. 2019, 8, 145–157. [Google Scholar] [CrossRef]
  79. Ji, S.; Torres, S.A.G.; Chen, J.; Lei, L.; Li, L. Molecular Dynamics Simulation of Film Water Thickness and Properties at Different Interfaces in Partially Saturated Frozen Soil Systems. Sci. Rep. 2025, 15, 2343. [Google Scholar] [CrossRef]
  80. Williams, P.L.; Mishin, Y.; Hamilton, J.C. An Embedded-Atom Potential for the Cu–Ag System. Model. Simul. Mat. Sci. Eng. 2006, 14, 817. [Google Scholar] [CrossRef]
  81. Heinz, H.; Vaia, R.A.; Farmer, B.L.; Naik, R.R. Accurate Simulation of Surfaces and Interfaces of Face-Centered Cubic Metals Using 12−6 and 9−6 Lennard-Jones Potentials. J. Phys. Chem. C 2008, 112, 17281–17290. [Google Scholar] [CrossRef]
  82. Mamatkulov, S.; Polák, J.; Razzokov, J.; Tomaník, L.; Slavíček, P.; Dzubiella, J.; Kanduč, M.; Heyda, J. Unveiling the Borohydride Ion through Force-Field Development. J. Chem. Theory Comput. 2024, 20, 1263–1273. [Google Scholar] [CrossRef]
Figure 1. (a) TEM image of Ag NPs derived from the NF extract. (b) High-resolution TEM image highlighting the lattice fringes of the crystalline Ag NPs. (c) Selected area electron diffraction (SAED) pattern confirming the crystalline nature of the synthesized Ag NPs. (d) Size distribution profile of Ag NP diameters corresponding to the particles in (a). (e) EDX spectrum from SEM-EDX analysis showing elemental composition of the Ag NPs. (f) FT-IR spectra comparing the NF extract (red line) and Ag NPs synthesized using the NF extract (blue line). Reprinted/adapted from ref. [12], Copyright (2025), with permission from Elsevier.
Figure 1. (a) TEM image of Ag NPs derived from the NF extract. (b) High-resolution TEM image highlighting the lattice fringes of the crystalline Ag NPs. (c) Selected area electron diffraction (SAED) pattern confirming the crystalline nature of the synthesized Ag NPs. (d) Size distribution profile of Ag NP diameters corresponding to the particles in (a). (e) EDX spectrum from SEM-EDX analysis showing elemental composition of the Ag NPs. (f) FT-IR spectra comparing the NF extract (red line) and Ag NPs synthesized using the NF extract (blue line). Reprinted/adapted from ref. [12], Copyright (2025), with permission from Elsevier.
Molecules 30 03738 g001
Figure 2. Methyl orange dye degradation in the presence of NaBH 4 without Ag NPs. After 30 min, the solution color virtually remained unchanged.
Figure 2. Methyl orange dye degradation in the presence of NaBH 4 without Ag NPs. After 30 min, the solution color virtually remained unchanged.
Molecules 30 03738 g002
Figure 3. (a) Illustration of the decolorization of a yellow MO solution by the synthesized Ag NPs in the presence of NaBH 4 . (b) UV–Vis spectral changes in the MO degradation by the synthesized Ag NPs at 0.05 M NaBH 4 . (c) UV–Vis spectral changes in the MO degradation by the synthesized Ag NPs at 0.20 M NaBH 4 . (d) Percent degradation of MO dye as a function of time at 0.05, 0.10, 0.15, 0.20, 0.25, and 0.30 M NaBH 4 , respectively. (e) Percent degradation of MO dye at the first 3.5 min as a function of NaBH 4 concentration. (f) Pseudo-first-order kinetics of MO degradation at different NaBH 4 concentrations.
Figure 3. (a) Illustration of the decolorization of a yellow MO solution by the synthesized Ag NPs in the presence of NaBH 4 . (b) UV–Vis spectral changes in the MO degradation by the synthesized Ag NPs at 0.05 M NaBH 4 . (c) UV–Vis spectral changes in the MO degradation by the synthesized Ag NPs at 0.20 M NaBH 4 . (d) Percent degradation of MO dye as a function of time at 0.05, 0.10, 0.15, 0.20, 0.25, and 0.30 M NaBH 4 , respectively. (e) Percent degradation of MO dye at the first 3.5 min as a function of NaBH 4 concentration. (f) Pseudo-first-order kinetics of MO degradation at different NaBH 4 concentrations.
Molecules 30 03738 g003
Figure 4. UV–Vis spectral changes in the MO degradation by the synthesized Ag NPs in the presence of 0.20 M NaBH 4 at (a) 25 μL, (b) 50 μL, (c) 75 μL, and (d) 100 μL of 0.20 M NaBH 4 , respectively. (e) Percent degradation of MO dye as a function of time at 25 μL, 50 μL, 75 μL, and 100 μL of 0.20 M NaBH 4 , respectively. (f) Corresponding pseudo-first-order kinetics of MO degradation at different NaBH 4 concentrations.
Figure 4. UV–Vis spectral changes in the MO degradation by the synthesized Ag NPs in the presence of 0.20 M NaBH 4 at (a) 25 μL, (b) 50 μL, (c) 75 μL, and (d) 100 μL of 0.20 M NaBH 4 , respectively. (e) Percent degradation of MO dye as a function of time at 25 μL, 50 μL, 75 μL, and 100 μL of 0.20 M NaBH 4 , respectively. (f) Corresponding pseudo-first-order kinetics of MO degradation at different NaBH 4 concentrations.
Molecules 30 03738 g004
Figure 5. (a) UV–Vis spectral changes in the degradation of the commercial synthetic dye by the synthesized Ag NPs. (b) Percent degradation of the commercial synthetic dye. As can be seen, the MO degradation reached 89% within only 5 min.
Figure 5. (a) UV–Vis spectral changes in the degradation of the commercial synthetic dye by the synthesized Ag NPs. (b) Percent degradation of the commercial synthetic dye. As can be seen, the MO degradation reached 89% within only 5 min.
Molecules 30 03738 g005
Figure 6. Snapshots from MD simulations over the course of 0 to 5.57 ns for the Ag–water interface model containing 2000 H 2 O molecules, 1 MO dye, and 1 Na BH 4 molecule, respectively. The time series is selected to illustrate the adsorption process of MO onto the Ag(111) surface. In parallel to the lateral perspective, we also show a top view of all atoms, except those of water, within 6 Å distance from the Ag(111) surface.
Figure 6. Snapshots from MD simulations over the course of 0 to 5.57 ns for the Ag–water interface model containing 2000 H 2 O molecules, 1 MO dye, and 1 Na BH 4 molecule, respectively. The time series is selected to illustrate the adsorption process of MO onto the Ag(111) surface. In parallel to the lateral perspective, we also show a top view of all atoms, except those of water, within 6 Å distance from the Ag(111) surface.
Molecules 30 03738 g006
Figure 7. (a) Snapshot from the MD simulation at 19.45 ns showing the adsorption of four BH 4 ions onto the Ag(111) surface. The right panel presents the top view of atoms, except H 2 O , within 6 Å above the Ag(111) surface. The numbers 1–4 denote the four BH 4 ions, while the orange rectangle outlines the simulation cell (periodic in x and y; free along z). (b) Normalized radial distribution functions, g(r), between Ag atoms and various interacting species: N atoms of the azo group in MO dye (blue solid curve), O atoms of water (red dotted curve), H atoms of BH 4 (green dashed curve), and Na + ions (black dash–dotted curve). The first peak maxima of g(r) for the Ag—N(–N=N–), Ag—O( H 2 O ), Ag—H( BH 4 ), and Ag— Na + pairs occur at 3.00 Å, 3.13 Å, 3.27 Å, and ~3.20 Å (barely observed), respectively. Data were sampled from the final 4 ns of the simulation. (c) Normalized g(r) for selected ion–dye atom pairs: Na + —O of the sulfonate group (black solid line), Na + —N of the azo group (cyan solid curve), and H( BH 4 )—N of the azo group (brown solid curve). The first peak maxima for the Na + —O( SO 3 ), Na + —N(–N=N–), and H( BH 4 )—N(–N=N–) pairs are located at ~4.70 Å and ~3.00 Å (the latter two barely observed). Data were taken from the same 4 ns runs as in (b).
Figure 7. (a) Snapshot from the MD simulation at 19.45 ns showing the adsorption of four BH 4 ions onto the Ag(111) surface. The right panel presents the top view of atoms, except H 2 O , within 6 Å above the Ag(111) surface. The numbers 1–4 denote the four BH 4 ions, while the orange rectangle outlines the simulation cell (periodic in x and y; free along z). (b) Normalized radial distribution functions, g(r), between Ag atoms and various interacting species: N atoms of the azo group in MO dye (blue solid curve), O atoms of water (red dotted curve), H atoms of BH 4 (green dashed curve), and Na + ions (black dash–dotted curve). The first peak maxima of g(r) for the Ag—N(–N=N–), Ag—O( H 2 O ), Ag—H( BH 4 ), and Ag— Na + pairs occur at 3.00 Å, 3.13 Å, 3.27 Å, and ~3.20 Å (barely observed), respectively. Data were sampled from the final 4 ns of the simulation. (c) Normalized g(r) for selected ion–dye atom pairs: Na + —O of the sulfonate group (black solid line), Na + —N of the azo group (cyan solid curve), and H( BH 4 )—N of the azo group (brown solid curve). The first peak maxima for the Na + —O( SO 3 ), Na + —N(–N=N–), and H( BH 4 )—N(–N=N–) pairs are located at ~4.70 Å and ~3.00 Å (the latter two barely observed). Data were taken from the same 4 ns runs as in (b).
Molecules 30 03738 g007
Figure 8. (a) Optimized geometry of a methyl orange dye with atom numbering. (b) A table of condensed Fukui index values ( f + and f ) of each atom of the optimized MO dye shown in (a). The darkest green color indicates the highest f + value, whilst the darkest orange color refers to the highest f value.
Figure 8. (a) Optimized geometry of a methyl orange dye with atom numbering. (b) A table of condensed Fukui index values ( f + and f ) of each atom of the optimized MO dye shown in (a). The darkest green color indicates the highest f + value, whilst the darkest orange color refers to the highest f value.
Molecules 30 03738 g008
Figure 9. (a) Relative free-energy profile (in kcal/mol) for the first step of MO degradation via (i) a direct hydride transfer (TS1) and (ii) Na + -mediated hydride transfer (TS2) from BH 4 to N9 atom of the azo group. The red curved arrows indicate electron flow. (b) Three-dimensional-optimized structures belong to those in (a), with distance labels in Å. Cartesian coordinates of all intermediates and transition states are provided in the Supporting Information.
Figure 9. (a) Relative free-energy profile (in kcal/mol) for the first step of MO degradation via (i) a direct hydride transfer (TS1) and (ii) Na + -mediated hydride transfer (TS2) from BH 4 to N9 atom of the azo group. The red curved arrows indicate electron flow. (b) Three-dimensional-optimized structures belong to those in (a), with distance labels in Å. Cartesian coordinates of all intermediates and transition states are provided in the Supporting Information.
Molecules 30 03738 g009
Figure 10. (a) Relative free-energy profile (in kcal/mol) for the second step of MO degradation. Since this step involves proton transfer processes, the relative free energies are corrected to pH values of 3 (red line), 5 (blue line), and 9 (green line), as described in the main text. This free-energy adjustment reveals the influence of pH on the energetics and mechanism of MO degradation. The red curved arrows indicate electron flow. (b) Three-dimensional-optimized structures belong to those in (a), with distance labels in Å. Cartesian coordinates of all intermediates and transition states are provided in the Supporting Information.
Figure 10. (a) Relative free-energy profile (in kcal/mol) for the second step of MO degradation. Since this step involves proton transfer processes, the relative free energies are corrected to pH values of 3 (red line), 5 (blue line), and 9 (green line), as described in the main text. This free-energy adjustment reveals the influence of pH on the energetics and mechanism of MO degradation. The red curved arrows indicate electron flow. (b) Three-dimensional-optimized structures belong to those in (a), with distance labels in Å. Cartesian coordinates of all intermediates and transition states are provided in the Supporting Information.
Molecules 30 03738 g010
Figure 11. Simulated UV–Vis spectra obtained from TD-DFT calculations. The spectra compare methyl orange (MO, orange line), structure P1 (sodium 4-aminobenzenesulfonate, solid black line), and structure P2 (N,N-dimethyl-p-phenylenediamine, dashed black line).
Figure 11. Simulated UV–Vis spectra obtained from TD-DFT calculations. The spectra compare methyl orange (MO, orange line), structure P1 (sodium 4-aminobenzenesulfonate, solid black line), and structure P2 (N,N-dimethyl-p-phenylenediamine, dashed black line).
Molecules 30 03738 g011
Table 1. Summary of the optimized parameters for the synthesis of Ag NPs using NF extract [12].
Table 1. Summary of the optimized parameters for the synthesis of Ag NPs using NF extract [12].
ConditionOptimized Parameter
pH9
Concentration of Ag NO 3 1.0 mM
Concentration of NF extract0.2% w/v
Volume ratio of NF extract: Ag NO 3 1:20
Temperature60 °C
Reaction time45 min
Table 2. A summary of the percent of MO degradation by the synthesized Ag NPs in the presence of 0.05, 0.10, 0.15, 0.20, 0.25, and 0.30 M NaBH 4 , along with the corresponding pseudo-first-order kinetics parameters.
Table 2. A summary of the percent of MO degradation by the synthesized Ag NPs in the presence of 0.05, 0.10, 0.15, 0.20, 0.25, and 0.30 M NaBH 4 , along with the corresponding pseudo-first-order kinetics parameters.
NaBH 4 , (M)% DegradationR2k (min−1)Time (min)
0.0595.990.85330.0309956.5
0.1095.550.94870.6868164.5
0.1595.330.97620.9293364.5
0.2099.500.99611.5026682.5
0.2599.660.98242.4571282.5
0.3099.030.98061.9120672.5
Table 3. A summary of the percent of MO degradation and pseudo-first-order kinetics parameters for the different volumes of Ag NP catalysts ranging from 25 μL to 100 μL (see Figure 4e,f).
Table 3. A summary of the percent of MO degradation and pseudo-first-order kinetics parameters for the different volumes of Ag NP catalysts ranging from 25 μL to 100 μL (see Figure 4e,f).
Volume of Ag NPs (μL)% DegradationR2k (min−1)Time (min)
2595.360.99290.5235834.5
5095.730.99631.4053661.5
7595.580.99252.4042231.0
10099.260.99052.7631691.0
Table 4. Comparison of the results of the catalytic MO degradation in the presence of NaBH 4 with other studies. All the collected studies presented only focus on the green synthesis of Ag NPs using biomolecules from natural extracts.
Table 4. Comparison of the results of the catalytic MO degradation in the presence of NaBH 4 with other studies. All the collected studies presented only focus on the green synthesis of Ag NPs using biomolecules from natural extracts.
Source of Ag NP SynthesisSize (nm) aDegradation Percentage (%)Reaction Rate (min−1)Time (min)[Ref.] (Year)
Trigonella foenum-graecum seeds10–30100 b0.66266[20] (2014)
Biophytum sensitivum19.06100 b0.27589[21] (2015)
Punica granatum36990.217512[22] (2015)
Anacardium occidentale testa25100 b0.117820[23] (2016)
Sterculia acuminata fruit10100 b0.08793[24] (2016)
Mussaenda erythrophylla leaf24–91--45[25] (2016)
Zanthoxylum armatum leaves15–50-0.00186>24 h[26] (2016)
Durio zibethinus10–25100 b0.6367.5[27] (2018)
Medicago polymorpha25–33970.3483[28] (2019)
Euphorbia geniculata leaf1797.28-30[29] (2019)
Centella asiatica30–5084.38-180[30] (2020)
Prunus mume (P. mume) fruit3099.960.078530[31] (2020)
Terminalia arjuna leaf10–5086.680.16614[32] (2020)
Calendula officinalis50–60>950.1810[33] (2020)
Dodonaea viscosa6096.20.29254[34] (2021)
Eucalyptus globulus fruit20–100>90%0.24710[35] (2021)
Sargassum serratifolium27.84-0.158016[36] (2021)
Simarouba glauca oil seed meal4.61100 b0.05540[37] (2021)
Bacillus cereus5–7.06-0.0976>25[38] (2021)
Heterotheca subaxillaris flower20–30100 b0.1211[39] (2022)
Neem5–1394.270.085735[40] (2022)
Rhus javanica, Rumex hastatus, and Callistemon viminalis55–6780–83-120[41] (2022)
Sargassum horneri22.72-0.226622[42] (2022)
Tea leaf and aloe vera leaf901000.50012[43] (2023)
Klebsiella pneumoniae22.25–47.9926.6-50[44] (2024)
Camellia sinensis74.85930.08740[45] (2024)
Usnea longissima (Lichenan)6.3100 b1.4812.67[46] (2024)
Trigonella foenum-graecum L. seed (HM 425)2883.630.041239[47] (2024)
S. costus root aqueous2272.880.0072135[48] (2024)
Cymbopogon citratus (lemongrass) 15–62.595.820.041390[49] (2025)
Ligustrum ovalifolium flower 50–100920.83448[50] (2025)
Morinda
citrifolia leaf
22.7298.60.9236[51] (2025)
Dipterocarpus retusus branch2.7699.10.069214[52] (2025)
Annona reticulata endophyte175.295-15[53] (2025)
Paullinia cupana Kunth leaf 39.33–126.296.42 c0.0946 c14 c[54] (2025)
Zingiber sianginensis19.49980.99183.67[55] (2025)
Nypa fruticans fruit husk4>992.763~1This study
a A single value indicates the reported average diameter size. b An exact value was not reported; however, we provide 100% degradation if the study stated a complete reduction of MO and the corresponding UV–Vis spectrum at the reduction time showed almost no absorption peak. c The results are based on the extract collected during the rainy season.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wonglakhon, T.; Chonsakon, A.; Nuengmatcha, P.; Ninwong, B.; Zahn, D.; Thepchuay, Y. Green Synthesized Silver Nanoparticles from Biowaste for Rapid Dye Degradation: Experimental Investigation and Computational Mechanistic Insights. Molecules 2025, 30, 3738. https://doi.org/10.3390/molecules30183738

AMA Style

Wonglakhon T, Chonsakon A, Nuengmatcha P, Ninwong B, Zahn D, Thepchuay Y. Green Synthesized Silver Nanoparticles from Biowaste for Rapid Dye Degradation: Experimental Investigation and Computational Mechanistic Insights. Molecules. 2025; 30(18):3738. https://doi.org/10.3390/molecules30183738

Chicago/Turabian Style

Wonglakhon, Tanakorn, Areeya Chonsakon, Prawit Nuengmatcha, Benjawan Ninwong, Dirk Zahn, and Yanisa Thepchuay. 2025. "Green Synthesized Silver Nanoparticles from Biowaste for Rapid Dye Degradation: Experimental Investigation and Computational Mechanistic Insights" Molecules 30, no. 18: 3738. https://doi.org/10.3390/molecules30183738

APA Style

Wonglakhon, T., Chonsakon, A., Nuengmatcha, P., Ninwong, B., Zahn, D., & Thepchuay, Y. (2025). Green Synthesized Silver Nanoparticles from Biowaste for Rapid Dye Degradation: Experimental Investigation and Computational Mechanistic Insights. Molecules, 30(18), 3738. https://doi.org/10.3390/molecules30183738

Article Metrics

Back to TopTop