Abstract: ABSTRACT: Title: A Method for Green Synthesis of Gold Nanoparticles, and Palladium Nanoparticles Using Sarcolobus carinatus The present disclosure proposes a method for green synthesis of gold nanoparticles, and palladium nanoparticles using sarcolobus carinatus. The development of plant-mediated gold (AuNPs), and palladium (PdNPs) for anti-inflammatory and therapeutic uses. These nanoparticles are synthesized using biological sources such as plants, fungi, and bacteria. Sarcolobus carinatus aqueous extract is utilized for the biosynthesis of AuNPs, and PdNPs, which exhibited SPR peaks at 545 nm, indicating successful nanoparticle formation. AuNPs remained stable for 194 days at room temperature, 120 days at 4°C, and 9 days at -20°C. PdNPs exhibited stability for 135 days at room temperature, 105 days at 4°C, and 11 days at -20°C.
Description:DESCRIPTION:
Field of the invention:
The present disclosure generally relates to the technical field of green nanotechnology, and in specific relates to development of plant-mediated gold (AuNPs), and palladium (PdNPs) for anti-inflammatory and therapeutic uses.
Background of the invention:
Nanotechnology is an emerging field of research, and nanobiotechnology is gaining significant interest due to its ability to develop nanosized materials from biological origins for various applications. Metal nanoparticles, particularly gold (AuNPs), and palladium (PdNPs) have been extensively studied for their unique properties, including biocompatibility, chemical stability, and low toxicity. These nanoparticles are widely applied in medicine, biology, chemistry, and physics, with increasing attention in biomedicine, pharmacy, and bioscience.
Despite the advantages of metal nanoparticles, conventional synthesis methods rely on chemical and physical processes that are expensive, environmentally hazardous, and involve toxic reagents. These methods often produce nanoparticles with unpredictable size distribution and stability, limiting their effectiveness in biomedical and catalytic applications. The need for eco-friendly, cost-effective, and biologically safe alternatives to synthesize metal nanoparticles remains a significant challenge in nanotechnology.
Several methods have been developed to synthesize gold, palladium, and bimetallic nanoparticles. Chemical reduction methods use strong reducing agents like sodium borohydride or hydrazine to produce nanoparticles. Physical methods employ laser ablation, sputtering, or electrochemical deposition to generate nanoparticles. Microbial and plant-based green synthesis use biological systems like bacteria, fungi, or plant extracts to synthesize nanoparticles in an eco-friendly manner.
However, chemical reduction method involves hazardous chemicals that pose environmental and health risks. The resulting nanoparticles may also contain residual toxic substances, limiting their biomedical applications. High-energy consumption and expensive equipment make these physical methods impractical for large-scale production. While promising, the microbial synthesis requires strict sterile conditions and extended processing times, making it less efficient. Although safer and eco-friendly, many plant-mediated methods lack optimization for rapid and controlled nanoparticle formation, leading to variable sizes and stability issues.
Therefore, there is a need for development of plant-mediated gold (AuNPs), and palladium (PdNPs) for anti-inflammatory and therapeutic uses. There is also a need for an eco-friendly, non-toxic biosynthesis of metal nanoparticles using Sarcolobus carinatus plant extracts.
Objectives of the invention:
The primary objective of the invention is to provide biosynthesis of gold (AuNPs), and palladium (PdNPs) using Sarcolobus carinatus that eliminates the need for toxic chemical reagents, making the process environmentally sustainable.
The other objective of the invention is to provide nanoparticles that are synthesized at room temperature with significantly reduced reaction times, improving process efficiency.
The other objective of the invention is to provide bioactive compounds such as alkaloids, flavonoids, tannins, and saponins enhances the stability and potential biomedical applications of the nanoparticles.
Another objective of the invention is to provide nanoparticles that exhibit high optical activity, making them suitable for biomedical imaging, diagnostics, and catalytic applications.
The other objective of the invention is to provide synthesized nanoparticles that are characterized by controlled shapes (spherical, triangular, and hexagonal) and small average particle sizes, enhancing their applicability in nano-medicine.
Summary of the invention:
The present disclosure proposes a method for green synthesis of gold nanoparticles, and palladium, and nanoparticles using sarcolobus carinatus. The following presents a simplified summary in order to provide a basic understanding of some aspects of the claimed subject matter. This summary is not an extensive overview. It is not intended to identify key/critical elements or to delineate the scope of the claimed subject matter. Its sole purpose is to present some concepts in a simplified form as a prelude to the more detailed description that is presented later.
In order to overcome the above deficiencies of the prior art, the present disclosure is to solve the technical problem to provide development of plant-mediated gold (AuNPs), and palladium (PdNPs) for anti-inflammatory and therapeutic uses.
According to an aspect, the invention provides a method for synthesizing nanoparticles (NPs). First, a fruit paste is boiled in water through a bain- marie method. The boiled fruit paste is then filtered to prepare an aqueous fruit extract of Sarcolobus carinatus. Next, the aqueous fruit extract is mixed with an aqueous solution at a ratio of 1:6 to obtain a nanoparticles mixture. The aqueous solution comprises at least one of chloroauric acid (HAuCl₄.3H₂O) at a concentration of 1 mM, and palladium chloride (PdCl₂) at a concentration of 2 mM. Next, the nanoparticles mixture is incubated at room temperature for a required time period to formulate the nanoparticles (NPs). In one embodiment, the nanoparticles (NPs) are formulated as an injectable, topical, or oral pharmaceutical composition for inflammatory conditions.
In one embodiment, the nanoparticles mixture of the aqueous fruit extract with the aqueous solution of HAuCl₄.3H₂O is incubated at room temperature for a time period of at least 5 hrs to obtain gold nanoparticles (AuNPs). In one embodiment, the nanoparticles mixture of the aqueous fruit extract with the aqueous solution of PdCl₂ is incubated at room temperature for a time period of at least 1 hr to obtain palladium nanoparticles (PdNPs).
In one embodiment, the synthesized AuNPs exhibit a surface plasmon resonance (SPR) peak at 545 nm. The AuNPs are synthesized at pH 4 to 12 and remain stable for at least 37 days at room temperature. In one embodiment, The AuNPs are separated by centrifugation at 10,000 rpm for 15 minutes and subsequently purified to obtain pellet.
Further, objects and advantages of the present invention will be apparent from a study of the following portion of the specification, the claims, and the attached drawings.
Detailed description of drawings:
The accompanying drawings, which are incorporated in and constitute a part of the specification, illustrate an embodiment of the invention, and, together with the description, explain the principles of the invention.
FIG. 1 illustrates a flowchart of a method for synthesizing nanoparticles (NPs), in accordance to an exemplary embodiment of the invention.
FIGs. 2A-2C illustrate graphs depicting visible inspection and effect of time on gold nanoparticles (SCF AuNPs), in accordance to an exemplary embodiment of the invention.
FIGs. 3A-3D illustrate graphs depicting effect of temperature on SCF AuNPs, effect of sarcolobus carinatus fruit extract concentration, effect of salt concentration on SCF AuNPs, and depicting effect of pH on SCF AuNPs, in accordance to an exemplary embodiment of the invention.
FIGs. 4A-4C illustrate graphs depicting visible inspection of palladium nanoparticles (PdNPs) at different ratios, and effect of time on PdNPs Formation, in accordance to an exemplary embodiment of the invention.
FIG. 5A illustrates a graph depicting effect of temperature on PdNPs synthesis, in accordance to an exemplary embodiment of the invention.
FIG. 5B illustrates a graph depicting effect of Sarcolobus carinatus fruit extract concentration on PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 5C illustrates a graph depicting effect of PdCl₂ salt concentration on PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 5D illustrates a graph depicting effect of pH on PdNPs synthesis, in accordance to an exemplary embodiment of the invention.
FIG. 6A illustrates a FTIR spectra of the aqueous fruit extract of Sarcolobus carinatus (SC fruit extract), in accordance to an exemplary embodiment of the invention.
FIG. 6B illustrates a FTIR spectra of the synthesized SCF AuNPs, in accordance to an exemplary embodiment of the invention.
FIG. 7 illustrates a Raman spectral for Raman spectroscopy analysis of the synthesized gold nanoparticles (SCF AuNPs), in accordance to an exemplary embodiment of the invention.
FIG. 8 illustrates a XRD pattern of the synthesized SCF AuNPs, in accordance to an exemplary embodiment of the invention.
FIG. 9 illustrates a field emission scanning electron microscopy (FESEM) image of AuNPs, in accordance to an exemplary embodiment of the invention.
FIG. 10 illustrates an energy-dispersive X-ray (EDX) spectrum of the synthesized AuNPs with a quantitative elemental analysis, in accordance to an exemplary embodiment of the invention.
FIG. 11A illustrates a High-Resolution Transmission Electron Microscopy (HR-TEM) image of the AuNPs, in accordance to an exemplary embodiment of the invention.
FIG. 11B illustrates a distribution histogram of the SCF AuNPs, in accordance to an exemplary embodiment of the invention.
FIG. 11C illustrates Selected Area Electron Diffraction (SAED) pattern of the SCF AuNPs, in accordance to an exemplary embodiment of the invention.
FIG. 12A illustrates a zeta potential distribution graph of the SCF AuNPs, in accordance to an exemplary embodiment of the invention.
FIG. 12B illustrates a dynamic light scattering (DLS) particle size distribution graph of the SCF AuNPs, in accordance to an exemplary embodiment of the invention.
FIG. 13A illustrates a FTIR spectra of the SCF fruit extract, in accordance to an exemplary embodiment of the invention.
FIG. 13B illustrates a FTIR spectra of the synthesized SCF PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 14 illustrates a Raman spectral graph corresponding to the Raman spectroscopy analysis of palladium nanoparticles synthesized using the SCF PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 15 illustrates a XRD pattern of the synthesized SCF PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 16 illustrates a field emission scanning electron microscopy (FESEM) image of PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 17 illustrates an energy-dispersive X-ray (EDX) spectrum of the synthesized PdNPs with a quantitative elemental analysis, in accordance to an exemplary embodiment of the invention.
FIG. 18A illustrates a HR-TEM image of the SCF PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 18B illustrates a distribution histogram of the SCF PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 18C illustrates SAED pattern of the SCF PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 19A illustrates a zeta potential distribution graph of the SCF PdNPs, in accordance to an exemplary embodiment of the invention.
FIG. 19B illustrates a dynamic light scattering (DLS) particle size distribution graph of the SCF PdNPs, in accordance to an exemplary embodiment of the invention.
FIGs. 20A - 20D illustrate comparative effects of various green-synthesized nanoparticles and extracts on % membrane stabilization and hemolysis inhibition, in accordance to an exemplary embodiment of the invention.
FIGs. 21A – 21B illustrate effect of green-synthesized nanoparticles and extracts derived from SCF on the inhibition of protein denaturation, which serves as an in vitro indicator of anti-inflammatory activity, in accordance to an exemplary embodiment of the invention.
FIGs. 22A-22B illustrate UV- Visible spectra showing reduction in peak intensity for methylene blue in the presence of NaBH₄, SCF PdNPs and SCF AuNPs (25 µg/ml) over time, in accordance to an exemplary embodiment of the invention.
FIGs. 22C-22D illustrate corresponding pseudo-first-order kinetic plots indicating the reaction rate constants for each catalyst, in accordance to an exemplary embodiment of the invention.
FIG. 23A illustrate a UV-Visible spectra of SCF AuNPs at a concentration of 25 μg, in accordance to an exemplary embodiment of the invention
FIG. 23B illustrate a UV-Visible spectra of SCF AuNPs at a concentration of 50 μg, in accordance to an exemplary embodiment of the invention.
FIG. 23C illustrate a UV-Visible spectra of SCF AuNPs at a concentration of 100 μg, in accordance to an exemplary embodiment of the invention.
FIGs. 23D-23F illustrate kinetic plots displaying the corresponding pseudo-first-order reaction rate constants, in accordance to an exemplary embodiment of the invention.
FIGs. 24A-24C illustrate UV-Visible spectra of SCF PdNPs (25 μg, 50 μg, 100 μg) depicting dye degradation over 72 hours, in accordance to an exemplary embodiment of the invention.
FIGs. 24D-24F illustrate kinetic plots showing pseudo-first-order rate constants, in accordance to an exemplary embodiment of the invention.
Detailed invention disclosure:
Various embodiments of the present invention will be described in reference to the accompanying drawings. Wherever possible, same or similar reference numerals are used in the drawings and the description to refer to the same or like parts or steps.
The present disclosure has been made with a view towards solving the problem with the prior art described above, and it is an object of the present invention to provide development of plant-mediated gold (AuNPs), and palladium (PdNPs) for anti-inflammatory and therapeutic uses.
According to an exemplary embodiment of the invention, FIG. 1 refers to a flowchart 100 of a method for synthesizing nanoparticles (NPs). At step 102, approximately 30 g of fresh Sarcolobus carinatus fruits are thoroughly washed twice with tap water, followed by a final rinse with milliQ water to remove any dust particles. The cleaned fruits are then ground into a fruit paste using a mortar and pestle. The fruit paste is transferred to a beaker containing at least 100 mL of milliQ water and boiled for a minimum of 15 minutes using a bain- marie method on a hot plate. The boiled fruit paste is first filtered twice through Whatman No.1 filter paper, followed by membrane filtration using a 0.45 µm pore size membrane to obtain an aqueous fruit extract of Sarcolobus carinatus (SCF). The aqueous fruit extract is stored in a refrigerator and used within 24 hrs.
At step 104, the aqueous fruit extract is mixed with an aqueous solution at a ratio of 1:6 to obtain a nanoparticles mixture. The aqueous solution comprises at least one of chloroauric acid (HAuCl₄.3H₂O) at a concentration of 1 mM, palladium chloride (PdCl₂) at a concentration of 2 mM. In one embodiment, the aqueous solution of HAuCl₄·3H₂O is prepared by dissolving 0.0196 g of HAuCl₄·3H₂O (1 mM) in 100 mL of distilled water.
In one embodiment, the aqueous solution of HAuCl₄.3H₂O is prepared by dissolving 0.0347 g of PdCl₂ (2 mM) in 100 mL of milliQ water. To aid dissolution, one drop of 0.1N HCl is added to the metal salt, followed by mixing with a glass rod for 1 minute. The solution is then placed on a magnetic stirrer at 150 rpm and stirred at room temperature for 1 hr.
At step 106, the nanoparticles mixture is incubated at room temperature for a required time period to formulate the nanoparticles (NPs). In one embodiment, the nanoparticles (NPs) are formulated as an injectable, topical, or oral pharmaceutical composition for inflammatory conditions.
In one embodiment, the nanoparticles mixture of the aqueous fruit extract with the aqueous solution of HAuCl₄.3H₂O is incubated at room temperature for a time period of at least 5 hrs to obtain gold nanoparticles (SCF AuNPs). In one embodiment, the nanoparticles mixture of the aqueous fruit extract with the aqueous solution of PdCl₂ is incubated at room temperature for a time period of at least 1 hr to obtain palladium nanoparticles (SCF PdNPs).
In one embodiment, the synthesized SCF AuNPs exhibit a surface plasmon resonance (SPR) peak at 545 nm. The SCF AuNPs are synthesized at a pH value varies between 4 and 12 and remain stable for at least 37 days at room temperature. In one embodiment, The SCF AuNPs are separated by centrifugation at 10,000 rpm for a time period of 15 min and subsequently purified to obtain pellet.
In one embodiment, the synthesized AuNPs may be referred to as SCF-AuNPs. Similarly, the synthesized PdNPs may be referred to as SCF-PdNPs.
According to another exemplary embodiment of the invention, FIGs. 2A – 2B refer to graphs (200, 202) depicting visible inspection of the AuNPs. The addition of Sarcolobus carinatus aqueous fruit extract (SCF) to a salt solution (HAuCl₄·3H₂O) causes a noticeable colour change from pale yellow to purple. This colour shift is a key visual indicator of the reduction of gold ions (Au³⁺ to Au⁰) and the subsequent formation of gold nanoparticles (AuNPs). The purple colour suggests the successful synthesis of nanoparticles through the bio-reduction process facilitated by phytochemicals present in Sarcolobus carinatus. The presence of polyphenols, flavonoids, or other reducing agents in the plant extract is likely responsible for reducing gold ions and stabilizing the synthesized AuNPs. The intensity of the colour could be correlated with the concentration and size of the nanoparticles formed.
According to another exemplary embodiment of the invention, FIG. 2C refer to graphs 204 depicting effect of time on AuNPs. A mixture of Sarcolobus carinatus extract and HAuCl₄·3H₂O (1:6 ratio) is maintained at a room temperature, and absorbance measurements are recorded at different time periods include 0hr to 30 days. After 2 hr, the reaction initiated, evidenced by the development of a pale lavender colour with an absorbance of 0.195 at 546 nm. As the reaction progressed, the colour deepened, and both absorbance and λmax increased steadily over six days, reaching a peak absorbance of 0.996. The gradual increase in absorbance and colour intensity suggests a continuous nucleation and growth process of nanoparticles. The initial lag phase (0–2h) before colour change indicates the time required for sufficient accumulation of reduced gold atoms before nanoparticle formation becomes visible. Over time, as more gold ions are reduced and nanoparticles grow, a shift in λmax is observed, which may indicate an increase in particle size or shape anisotropy. The stability of the nanoparticles over six days suggests effective capping by biomolecules from the plant extract.
According to another exemplary embodiment of the invention, FIG. 3A refer to a graph 300 depicting effect of temperature on SCF AuNPs. The reaction is carried out at different temperatures varies between 40°C and 90°C. A slight variation in absorbance is observed at 90°C while maintaining a constant λmax of 535 nm. The absorbance increased from 0.497 to 0.574 with increasing temperature, indicating a faster reaction rate. Higher temperatures likely enhanced the reaction kinetics, leading to a faster reduction of gold ions and quicker nanoparticle formation. The constant λmax at 535 nm suggests that temperature changes did not significantly alter nanoparticle size or shape distribution, implying a uniform synthesis process. The increase in absorbance at higher temperatures suggests higher nanoparticle yield due to increased reaction efficiency. However, excessively high temperatures could lead to particle aggregation or non-uniform growth, requiring optimization for controlled synthesis.
According to another exemplary embodiment of the invention, FIG. 3B refers to a graph 302 depicting effect of sarcolobus carinatus fruit extract concentration. With 0.1ml-0.5ml (sarcolobus carinatus) fruit Extract. The colour and absorbance are increased from 0.196-0.306 with λmax 539-540nm. While AuNPs of P.ligularis fruit extract are synthesised with 1ml, 2ml, and 3ml at room temperature (25°C) for around 6h.
According to another exemplary embodiment of the invention, FIG. 3C illustrates a graph 304 depicting the effect of salt concentration on SCF AuNPs. As shown in FIG. 3C, an increase in salt concentration from 1 mM to 5 mM results in a visible color change from purple to blackish-purple. The colour and absorption intensities increased with increased concentrations of salt along with increased λmax from 545- 609nm. Studies of P. ligularis also reported synthesis of AuNPs with HAuCl4.3H2O volume (1ml, 2ml, 3ml, 4ml, 5ml, 6ml, and 7ml) for 6 hrs at room temperature.
According to another exemplary embodiment of the invention, FIG. 3D refers to a graph 306 depicting effect of the pH value on AuNPs. At pH 4–9, the maximum absorbance wavelength (λmax) initially increased from 538 nm to 573 nm within 1 hr, followed by a gradual decrease over 5 hr from 535 nm to 560 nm. At pH 10–12, the initial λmax decreased from 549 nm to 524 nm within 1 hr but then increased gradually over 5 hr from 551 nm to 528 nm. The pH value influences the size and morphology of the nanoparticles due to its impact on the ionization of biomolecules in the plant extract. The shift in λmax suggests changes in nanoparticle size and shape with the varying pH value. At lower pH value (acidic conditions), larger nanoparticles or aggregation may occur, leading to red shifts in λmax. At higher pH value (alkaline conditions), the shift toward lower λmax suggests smaller nanoparticles, which are typically more stable due to enhanced electrostatic repulsion preventing aggregation. The observed variations in absorbance and wavelength shifts indicate that pH plays a crucial role in controlling nanoparticle size distribution and stability.
According to another exemplary embodiment of the invention, FIGs. 4A - 4B refer to visible inspection graphs (400, 402) of PdNPs at different ratios. The addition of Sarcolobus carinatus aqueous fruit extract (SCF) to a 2 mM PdCl₂ solution results in a colour change from pale yellow to brown, indicating the reduction of palladium ions and the formation of palladium nanoparticles (PdNPs). Different extract-to-salt ratios (1:1 to 1:14) are tested, with colour intensity varying based on the ratio used. The development of brown coloration suggests the reduction of Pd²⁺ ions to Pd⁰ nanoparticles by phytochemicals in the plant extract. UV- Visible spectrophotometric analysis is conducted in the 300–700 nm range to measure colour intensity variations.
In one embodiment, FIG. 4B shows SCF PdNPs synthesis at different extract-to-salt ratios (1:1–1:14) at room temperature, confirming that nanoparticle formation occurs within this range. The intensity of brown colour increases with certain ratios, indicating an optimal balance between reducing agents and metal ions for efficient nanoparticle synthesis. In one embodiment, FIG. 3A highlights the selected optimal ratio (1:6), which demonstrated effective SCF PdNPs formation at room temperature within 1 hr of incubation. The choice of the 1:6 ratio suggests that it provides an ideal concentration of reducing agents for stable nanoparticle formation without excessive aggregation or incomplete reduction.
In one embodiment, the reaction mixture (1:6 ratio of plant extract to PdCl₂) is kept at a room temperature, and absorbance measurements are taken at 1-hr intervals from 0 to 24 hr. After 1 hr, the reaction initiated, producing a pale brown colour. Over time, the brown colour deepened, indicating progressive SCF PdNPs formation and an increase in nanoparticle concentration. The increase in colour intensity suggests a continuous reduction of Pd²⁺ ions and nanoparticle growth over time. The formation process appears gradual, with no immediate saturation observed within the first 24 hr. The UV- Visible absorbance trend likely shows an increase in peak intensity over time, confirming the progressive formation of PdNPs. The 1-hr initiation period indicates the time required for the plant extract's biomolecules to begin effective reduction of Pd²⁺ ions.
According to another exemplary embodiment of the invention, FIG. 4C refers to a graph 404 depicting effect of time on SCF PdNPs stability (Long-Term Study). The same reaction mixture is monitored for up to 8 days to study nanoparticle stability and continued formation. Colour intensity continued to increase, indicating ongoing reduction and potential particle maturation over several days. The UV-V is spectrum likely showed shifts in peak intensity, confirming sustained SCF PdNPs formation over an extended period. The gradual increase in colour intensity over 8 days suggests a longer-term reduction process or possible nanoparticle aggregation over time. Stability over multiple days indicates efficient capping and stabilization by biomolecules in the plant extract, preventing rapid aggregation or precipitation of PdNPs. The extended reaction time might contribute to size tuning, as nanoparticle growth could continue beyond the initial incubation period. The findings suggest that SCF PdNPs synthesized under these conditions remain stable for at least 8 days, which is important for further applications in catalysis, biomedicine, or material science.
According to another exemplary embodiment of the invention, FIG. 5A refers to a graph 500 depicting effect of temperature on SCF PdNPs synthesis. The reaction is conducted at different temperatures ranging from 40°C to 90°C. A slight variation in absorbance is observed at a temperature of 90°C, but the colour remained constant, indicating minimal impact on the visual characteristics of PdNPs formation. Temperature plays a role in reaction kinetics, with higher temperatures typically accelerating the reduction of Pd²⁺ to Pd⁰ nanoparticles. The constant colour despite increasing temperature suggests that temperature changes did not significantly affect particle size or aggregation, possibly due to strong capping by biomolecules in the plant extract. The slight variation in absorbance at 90°C could indicate minor changes in nanoparticle growth rate but not significant alteration in particle distribution. This stability across temperatures suggests that PdNPs formation is not highly temperature-dependent beyond a certain threshold.
According to another exemplary embodiment of the invention, FIG. 5B refers to a graph 502 depicting effect of Sarcolobus carinatus fruit extract concentration on PdNPs. The reaction is carried out using different concentrations of Sarcolobus carinatus fruit extract ranging from 0.1 mL to 0.5 mL. The colour intensity increased with higher extract concentrations. Higher concentrations of plant extract provide more reducing and stabilizing biomolecules, leading to faster and more efficient Pd²⁺ reduction. Increased extract concentration might also enhance capping, leading to better nanoparticle stabilization and preventing excessive aggregation. The increase in colour intensity suggests that more nanoparticles are formed as extract concentration increases, potentially leading to higher yield. However, excessive extract concentration beyond an optimal range might result in over-capping, which can affect nanoparticle size and properties.
According to another exemplary embodiment of the invention, FIG. 5C refers to a graph 504 depicting effect of PdCl₂ salt concentration on PdNPs. The reaction is carried out using varying PdCl₂ concentrations (1 mM to 5 mM). The colour changed from brown to black as PdCl₂ concentration increased. Higher concentrations resulted in increased colour intensity. The transition from brown to black suggests that higher salt concentrations lead to the formation of larger nanoparticles or nanoparticle aggregation. Increased precursor concentration (Pd²⁺ ions) allows for higher nanoparticle production, increasing overall absorbance and colour intensity. At very high concentrations, uncontrolled nucleation might lead to poly-dispersed particles or aggregation, which may explain the black colour development. This indicates that while higher salt concentrations favour nanoparticle formation, optimal conditions are needed to control particle size and stability.
According to another exemplary embodiment of the invention, FIG. 5D refers to a graph 506 depicting effect of pH on SCF PdNPs synthesis. The reaction is performed across a pH range of 4 to 12. At pH values 4 to 9, colour intensity increased within the first hour. At pH values 10 to 12, colour intensity increased further, with a significant rise in Surface Plasmon Resonance (SPR) peak observed at pH values 11 and 12. The pH values influence the reduction rate, nanoparticle size, and stability by affecting the ionization of plant extract biomolecules. At acidic pH values (4 to 9), nanoparticles are formed, but the reduction rate is slower, leading to smaller particles with moderate colour intensity. At alkaline pH values (10 to 12), nanoparticle formation is enhanced due to the faster reduction of Pd²⁺ ions, resulting in increased SPR peak intensity at the pH values 11 and 12. The shift in SPR peak position indicates changes in nanoparticle size and shape, suggesting larger or anisotropic particles at higher pH values. Optimal pH value selection is crucial to balance nanoparticle formation, stability, and desired size distribution.
According to another exemplary embodiment of the invention, FIG. 6A refers to a FTIR spectra 600 of the aqueous fruit extract of Sarcolobus carinatus (SC fruit extract). FIG. 6B refers to a FTIR spectra 602 of the synthesized AuNPs. Phytochemical screening of the SCF extract revealed the presence of a diverse array of biomolecules, including alkaloids, carbohydrates, cardiac glycosides, coumarins, flavonoids, glycosides, proteins, phytosteroids, quinones, and reducing sugars. FTIR analysis is conducted to identify the principal functional groups within the SC fruit extract that are responsible for the reduction and stabilization (capping) of gold nanoparticles (SCF AuNPs).
Referring to FIGs. 6A – 6B, a comparative analysis of the FTIR spectra of the SCF fruit extract and the synthesized AuNPs demonstrates significant shifts in characteristic absorption peaks, indicating chemical interactions and involvement of specific functional groups in the nanoparticle formation process. Notably, absorption peaks observed at 3971 cm⁻¹ and 3919 cm⁻¹ in the fruit extract are shifted to 3972 cm⁻¹ in the AuNPs spectrum, corresponding to O–H stretching vibrations characteristic of alcohol groups. Additionally, the peaks at 3694 cm⁻¹, 3663 cm⁻¹, 3624 cm⁻¹, and 3601 cm⁻¹ in the extract shifted to a single band at 3665 cm⁻¹ in the AuNPs spectrum, which may be attributed to O–H stretching in phenolic compounds.
The alkyne-associated peak at 2037 cm⁻¹ in the extract shifted to 2023 cm⁻¹ in the nanoparticle spectrum, indicative of C≡C triple bond vibrations. The prominent absorption at 1639 cm⁻¹ in the extract shifted to 1632 cm⁻¹ in SCF AuNPs, corresponding to stretching vibrations of C=O (carbonyl) and N–H bonds. Peaks at 1639 cm⁻¹, 1463 cm⁻¹, 1425 cm⁻¹, and 1408 cm⁻¹ are absent in the SCF AuNP spectrum, suggesting their active participation in the reduction and formation of nanoparticles.
Further, the peak at 1062 cm⁻¹ in the extract, associated with C–C stretching in alcohols, is shifted to 1017 cm⁻¹ in the AuNPs. Peaks at 677 cm⁻¹, 553 cm⁻¹, and 521 cm⁻¹ in the extract are shifted to 951 cm⁻¹, 481 cm⁻¹, 443 cm⁻¹, and 430 cm⁻¹ in the AuNPs spectrum, indicative of C=C stretching in aromatic compounds. Broad peaks observed between 3390 cm⁻¹ and 3102 cm⁻¹ in the extract are absent in the SCF AuNP spectrum, further suggesting the involvement of these functional groups in the reduction and stabilization of gold nanoparticles. These spectral changes collectively confirm the role of various phytoconstituents in the SCF extract as reducing and capping agents in the green synthesis of AuNPs.
According to another exemplary embodiment of the invention, FIG. 7 refers to a Raman spectral 700 for Raman spectroscopy analysis of the synthesized gold nanoparticles (AuNPs). The Raman spectral 700 highlighting key vibrational modes associated with various functional groups. Prominent peaks observed at 1594 cm⁻¹ and 1329 cm⁻¹ are attributed to the presence of nitro-substituted aromatic rings and strong aromatic ring vibrations, respectively. Additional medium-intensity bands are detected at 1801 cm⁻¹, 2043 cm⁻¹, and 2439 cm⁻¹, corresponding to vibrational modes of anhydride groups, isothiocyanate moieties, and weak P–H bonds, respectively. Low-intensity peaks at 1010 cm⁻¹ and 1196 cm⁻¹ are associated with aromatic ring stretching and very weak sulphonic acid group vibrations. Furthermore, moderate peaks are observed at 304 cm⁻¹, 588 cm⁻¹, 767 cm⁻¹, and 878 cm⁻¹, which are characteristic of metal–oxygen (M–O) bonds, carbon–bromine (C–Br) stretching, carbon–chlorine (C–Cl) stretching, carbon–sulfur (C–S) bonds, and weak ether (C–O–C) linkages, respectively. These Raman spectral 700 features confirm the involvement of various functional groups in the surface chemistry and stabilization of the AuNPs.
According to another exemplary embodiment of the invention, FIG. 8 refers to a XRD pattern 800 of the synthesized AuNPs. The XRD pattern 800 of the synthesized SCF AuNPs exhibited distinct diffraction peaks at 2θ values of 38.19°, 44.35°, 64.7°, 77.7°, and 81.76°. These peaks correspond to the (111), (200), (220), (311), and (222) crystallographic planes, respectively, of a face-centered cubic (FCC) structure of metallic gold. The diffraction pattern is in good agreement with the standard reference pattern from JCPDS card no. 04-0784, confirming the crystalline nature and phase purity of the synthesized SCF AuNPs.
Among the observed peaks, the (111) reflection was the most intense, indicating that the (111) plane is the preferred orientation during nanoparticle formation. The average crystallite size of the SCF AuNPs, calculated using the Scherrer equation, was approximately 8.34 nm. Similar diffraction patterns were reported for gold nanoparticles synthesized using other biological sources such as Dalbergia coromandeliana and Padina gymnospora. In contrast, nanoparticles synthesized from Passiflora ligularis exhibited a comparable FCC structure with a crystallite size of approximately 8.13 nm.
According to another exemplary embodiment of the invention, FIG. 9 refers to a field emission scanning electron microscopy (FESEM) image 900 of AuNPs. The FESEM analysis confirmed that the synthesized AuNPs exhibit a predominantly spherical morphology, with particle sizes varies between 20 nm and 41 nm, as observed at a 100 nm scale and an accelerating voltage of 7 kV. Additionally, energy-dispersive X-ray (EDX) spectroscopy revealed a dominant peak at 5 keV, confirming gold (Au) as the principal element, comprising 84.73% of the sample. Other detected elements included carbon (8.19%), oxygen (6.4%), and trace amounts of chlorine (Cl). These results are consistent with AuNPs synthesized using Nostoc calcicola, which also demonstrated a spherical shape with a broader size distribution varies betweeen 20 nm and 140 nm.
In one embodiment, high-resolution TEM imaging 900 revealed that SCF-AuNPs exhibited diverse morphologies, including spherical, triangular, and hexagonal shapes. Image analysis conducted using ImageJ software, along with the corresponding size distribution histogram, showed particle sizes ranging from 5 nm to 45 nm, with an average diameter of 13.17 nm. These findings are consistent with AuNPs synthesized from Terminalia chebula, which demonstrated a broader size distribution of 6–60 nm. Furthermore, the Selected Area Electron Diffraction (SAED) pattern exhibited bright circular rings, which were indexed using Miller indices and correlated with standard gold diffraction planes. The observed diffraction planes—(111), (200), (220), and (311)—confirmed that the synthesized AuNPs possess a face-centered cubic (FCC) crystalline structure characteristic of metallic gold.
In one embodiment, the Zeta potential distribution showed a negative surface charge of -19.6 mV, closely matching the -16.4 mV value reported for Nostoc calcicola-derived SCF AuNPs. This negative charge is attributed to surface-bound reducing agents from the plant extract, including phytochemicals such as amino acids and flavonoids, which play a role in the capping and stabilization of SCF AuNPs. In one embodiment, dynamic light scattering (DLS) analysis revealed that AuNPs are poly-disperse, with a Z-average size of 138 nm and a poly-dispersity index (PDI) of 33.2%, indicating some variation in particle size distribution. In one embodiment, the SCF extract contains a variety of bioactive compounds, including alkaloids, carbohydrates, cardiac glycosides, coumarins, flavonoids, glycosides, proteins, phytosteroids, quinones, and reducing sugars.
According to another exemplary embodiment of the invention, FIG. 10 refers to an energy-dispersive X-ray (EDX) spectrum 1000 of the synthesized AuNPs and a quantitative elemental analysis derived from the EDX spectrum 1000. The EDX spectrum 1000 exhibits a prominent peak at approximately 5 keV, which is characteristic of the gold (Au) Lα emission line, thereby confirming gold as the predominant element in the sample. Quantitative analysis indicates that gold constitutes 84.73 wt% of the sample, as shown in FIG. 18B. In addition to gold, the spectrum revealed the presence of carbon (8.19 wt%), oxygen (6.4 wt%), and trace amounts of chlorine (Cl), likely originating from residual organic capping agents or precursors used during synthesis. These compositional results are consistent with previously reported biosynthesized SCF AuNPs, such as those synthesized using Nostoc calcicola, which also demonstrated spherical morphology and a broad particle size distribution ranging from 20 nm to 140 nm.
According to another exemplary embodiment of the invention, FIG. 11A refers to a High-Resolution Transmission Electron Microscopy (HR-TEM) image 1100 of the SCF AuNPs at 20 nm. FIG. 11B refers to a distribution histogram 1102 of the SCF AuNPs. FIG. 11C refers to Selected Area Electron Diffraction (SAED) pattern 1104 of the SCF AuNPs.
FIG. 11A presents HR-TEM images 1100 of the SCF AuNPs, revealing the presence of nanoparticles with spherical, triangular, and hexagonal morphologies. Image analysis was performed using ImageJ software, and the size distribution histogram shown in FIG. 11B indicates that the SCF AuNPs exhibited a size range of approximately 5–45 nm, with an average particle size of 13.17 nm. For comparison, gold nanoparticles synthesized using Terminalia chebula have been reported to exhibit an average particle size in the range of 6–60 nm.
FIG. 11C illustrates the selected area electron diffraction (SAED) pattern 1104 of the SCF AuNPs. The SAED pattern displays distinct and bright circular rings, which were indexed using Miller indices and found to correspond to the standard diffraction planes of crystalline gold. The rings were assigned to the (111), (200), (220), and (311) planes, confirming that the synthesized SCF AuNPs possess a face-centered cubic (FCC) crystalline structure.
According to another exemplary embodiment of the invention, FIG. 12A refers to a zeta potential distribution graph 1200 of the SCF AuNPs. FIG. 12B refers to a dynamic light scattering (DLS) particle size distribution graph 1202 of the AuNPs.
The zeta potential of the SCF AuNPs is measured to be –19.6 mV, indicating moderate colloidal stability. This value is in close agreement with zeta potential values previously reported for biologically synthesized nanoparticles, such as those derived from Nostoc calcicola (–16.4 mV). The observed negative surface charge is attributed to surface-bound reducing and capping agents originating from phytochemicals in the plant extract, such as amino acids, flavonoids, and other bioactive compounds, which contribute to nanoparticle stabilization through electrostatic repulsion.
The DLS analysis (FIG. 12B) revealed that the SCF AuNPs exhibit a polydisperse size distribution with a Z-average particle size of 138 nm and a polydispersity index (PDI) of 0.332. The relatively broad distribution suggests the presence of a mixture of nanoparticle sizes, likely due to variation in phytochemical-mediated nucleation and growth during synthesis.
Table 1:
Size[nm] Area[%] SD[nm]
Peak1 1057.8 44.30 313.9
Peak2 138.30 52.68 62.60
Peak3 475.4 4.21 62.55
Hydrodynamic diameter 287.0nm
Polydispersity index 33.3%
According to another exemplary embodiment of the invention, FIG. 13A refers to a FTIR spectra 1300 of the SCF fruit extract. FIG. 13B refers to a FTIR spectra 1302 of the synthesized SCF PdNPs. The SCF fruit extract contains a complex mixture of phytochemicals, including alkaloids, carbohydrates, cardiac glycosides, coumarins, flavonoids, glycosides, proteins, phytosteroids, quinones, and reducing sugars. These biomolecules are believed to play critical roles as both reducing and capping agents during the synthesis SCF PdNPs.
A comparative analysis of the spectra reveals several key shifts in peak positions, indicative of chemical interactions between the phytochemicals and the palladium ions. The O–H stretching vibrations associated with alcohol groups in the extract, originally observed at 3892 cm⁻¹, were shifted to 3895 cm⁻¹ in the SCF PdNPs, while the peak at 3878 cm⁻¹ shifted to 3847 cm⁻¹, suggesting the presence of phenolic O–H groups. A peak at 3770 cm⁻¹ shifted to 3796 cm⁻¹, confirming the presence of alcoholic O–H stretching. Peaks at 3694 cm⁻¹, 3624 cm⁻¹, and 3562 cm⁻¹ in the extract were shifted to 3667 cm⁻¹, 3622 cm⁻¹, and 3562 cm⁻¹ in the SCF PdNPs, indicating strong O–H stretching vibrations from alcohol groups. Peaks at 3390 cm⁻¹, 3336 cm⁻¹, 3233 cm⁻¹, and 3225 cm⁻¹ in the extract shifted to 3448 cm⁻¹, 3403 cm⁻¹, 3326 cm⁻¹, and 3207 cm⁻¹ in the SCF PdNPs, indicating the presence of N–H stretching vibrations attributed to amide groups. The peak at 2037 cm⁻¹, corresponding to C≡C alkyne triple bonds, shifted to 2019 cm⁻¹ in the SCF PdNPs.
The amide-related vibrations observed at 1639 cm⁻¹ in the extract shifted slightly to 1642 cm⁻¹ in the PdNPs, confirming the involvement of C=O and N–H bond vibrations. The C–C stretching vibrations of alcohols, originally observed at 1062 cm⁻¹, were shifted to 1018 cm⁻¹. Aromatic C=C stretching vibrations, originally located at 677, 553, 521, and 429 cm⁻¹, were shifted and merged into a peak at 951 cm⁻¹, confirming the presence of aromatic ring structures.
According to another exemplary embodiment of the invention, FIG. 14 refers to a Raman spectral graph 1400 corresponding to the Raman spectroscopy analysis of palladium nanoparticles synthesized using the SCF PdNPs. The Raman spectrum confirms the presence of various functional groups and molecular vibrations associated with phytochemical residues adsorbed on the nanoparticle surface. Prominent peaks were observed at 1737 cm⁻¹ and 2331 cm⁻¹, which are attributed to the presence of lactone groups and P–H bonds, respectively. Medium intensity bands at 2052 cm⁻¹, 2164 cm⁻¹, and 2441 cm⁻¹ correspond to azide groups (N₃⁻), isothiocyanates (–N=C=S), and thiol-related vibrations (–SH), indicating the likely contribution of sulfur- and nitrogen-containing compounds in nanoparticle stabilization. Lower intensity peaks observed at 78.37 cm⁻¹, 265.21 cm⁻¹, 484 cm⁻¹, 633 cm⁻¹, and 846 cm⁻¹ are associated with C–C vibrations in aliphatic chains, strong Si–O–Si stretching, and weak C–O–C ether linkages. These spectral features suggest that various organic functional groups from the SCF extract interact with the PdNP surface and are likely involved in both reduction of palladium ions and surface stabilization of the nanoparticles.
According to another exemplary embodiment of the invention, FIG. 15 refers to a XRD pattern 1500 of the synthesized SCF PdNPs. The diffractogram displays distinct diffraction peaks at 2θ values of 38.19°, 44.35°, 64.7°, 77.7°, and 81.76°, which can be indexed to the (111), (200), (220), and (311) crystallographic planes, respectively, of a face-centered cubic (FCC) crystalline structure of metallic palladium, consistent with the JCPDS reference pattern No. 05-0681.
The (111) reflection was found to exhibit the highest intensity, suggesting it to be the predominant growth orientation and confirming the crystalline nature of the synthesized nanoparticles. Additional minor peaks corresponding to planes such as (100), (204), and (303) were also observed. These may be attributed to the presence of residual organic compounds from the SCF plant extract that participated in the synthesis and stabilization of the SCF PdNPs.
The average crystallite size of the SCF-PdNPs was estimated using the Scherrer equation and found to be approximately 13.67 nm. For comparison, palladium nanoparticles synthesized using Ananas comosus peel extract were reported to have an average crystallite size of 19.8 nm, also exhibiting a face-centered cubic structure.
According to another exemplary embodiment of the invention, FIG. 16 refers to a field emission scanning electron microscopy (FESEM) image 1600 of SCF PdNP. FESEM analysis confirmed that PdNPs are predominantly spherical in shape, as observed in images taken at 100 nm resolution with a 5 kV accelerating voltage. The particle size ranged between 20 nm and 41 nm. In one embodiment, energy dispersive X-ray (EDX) spectroscopy analysis further confirmed the elemental composition of the synthesized PdNPs, with a dominant peak at 5.417 keV, indicating that palladium is the major constituent, contributing 49.82% of the composition. Other detected elements included carbon (28.25%), oxygen (18.96%), and trace amounts of chlorine (Cl). The results are in agreement with those observed for PdNPs synthesized using Curcuma longa, which exhibited spherical morphology with sizes ranging from 40 nm to 90 nm. The EDX spectrum confirmed that the synthesized PdNPs are primarily composed of palladium, with minimal interference from other elements.
In one embodiment, high-resolution transmission electron microscopy (HR-TEM) images confirmed that the SCF PdNPs are spherical in shape. Image analysis using ImageJ software and the corresponding size distribution histogram revealed that the nanoparticles ranged in size from 5 nm to 25 nm, with an average particle size of 11.9 nm. Similar results are observed for PdNPs synthesized using Gymnema sylvestre, which had an average size of 10–20 nm. In one embodiment, selected area electron diffraction (SAED) pattern analysis displayed bright circular rings, which are indexed using Miller indices and correlated with standard palladium diffraction planes. The diffraction peaks corresponding to the (111), (200), (220), and (311) planes confirmed that the synthesized SCF PdNPs adopted a face-centered cubic (FCC) palladium structure.
In one embodiment, the zeta potential distribution of SCF PdNPs is found to be -10.3 mV, which is in close agreement with PdNPs synthesized using grape seed extract, which exhibited a zeta potential of –24 ± 1.21 mV. The negative charge on the nanoparticles is attributed to surface-bound reducing agents from the plant extract, with phytochemicals such as amino acids and flavonoids potentially playing a role in capping and stabilization of the PdNPs. In one embodiment, dynamic light scattering (DLS) analysis revealed that the PdNPs are poly-dispersed, with a Z-average size of 138 nm and a poly-dispersity index (PDI) of 55.5%.
According to another exemplary embodiment of the invention, FIG. 17 refers to an energy-dispersive X-ray (EDX) spectrum 1700 of the synthesized SCF PdNPs, and a quantitative elemental analysis derived from the EDX spectrum 1700. The EDX spectrum 1700 shows a prominent peak at approximately 5.417 keV, confirming the presence of palladium (Pd) as the major elemental constituent, accounting for approximately 49.82% of the total composition. In addition to palladium, the EDX analysis revealed the presence of carbon (28.25%), oxygen (18.96%), and trace amounts of chlorine (Cl). The presence of carbon and oxygen is attributed to phytochemicals from the SCF extract that may have acted as capping and stabilizing agents during nanoparticle formation. These findings corroborate with previous reports on Curcuma longa–mediated palladium nanoparticles, which were observed to be predominantly spherical in morphology with size distributions ranging between 40 nm and 90 nm. The EDX data thus confirms that the synthesized SCF PdNPs are predominantly composed of palladium, with associated organic residues from the plant matrix.
According to another exemplary embodiment of the invention, FIG. 18A refers to a HR-TEM image 1800 of the PdNPs. FIG. 18B refers to a distribution histogram 1802 of the PdNPs. FIG. 18C refers to SAED pattern 1804 of the SCF PdNPs.
The images 1800 confirm that the nanoparticles are predominantly spherical in morphology. Analysis of the TEM micrographs using ImageJ software yielded a particle size distribution histogram (FIG. 18B), indicating that the PdNPs ranged in size from 5 nm to 25 nm, with an average particle diameter of approximately 11.9 nm. These results are consistent with prior reports, such as those involving Gymnema sylvestre-mediated PdNPs, which exhibited size distributions in the range of 10–20 nm.
The SAED pattern 1802 of the SCF-PdNPs, shown in FIG. 18C, displays bright circular diffraction rings that were indexed using Miller indices. The observed diffraction rings correspond to the (111), (200), (220), and (311) crystallographic planes, consistent with a face-centered cubic (FCC) lattice structure of crystalline palladium, thereby corroborating the XRD findings.
According to another exemplary embodiment of the invention, FIG. 19A refers to a zeta potential distribution graph 1900 of the PdNPs. FIG. 19B refers to a dynamic light scattering (DLS) particle size distribution graph 1902 of the SCF PdNPs.
Referring to FIG. 19A, the measured zeta potential was −10.3 mV, indicating a moderately stable colloidal dispersion. Electrophoretic mobility -0.8039µm*cm/Vs. The observed negative surface charge is attributed to phytochemicals such as amino acids, flavonoids, and other functional biomolecules present in the SCF extract, which likely serve as reducing and capping agents during nanoparticle synthesis. These findings are in agreement with prior reports, such as grape seed extract–mediated PdNPs, which exhibited a zeta potential of approximately −24 ± 1.21 mV.
FIG. 19B presents the results of dynamic light scattering (DLS) analysis, which revealed that the SCF-PdNPs were polydisperse. The Z-average hydrodynamic diameter was measured to be 138 nm, with a polydispersity index (PDI) of 0.555, indicating a broad size distribution and the presence of some aggregation or variability in nanoparticle size.
Table 2:
Size[nm] Area [%] SD [nm]
Peak1 502.1 100.00 80.73
Peak2 - - -
Peak3 - - -
Hydrodynamic diameter 619.9nm
Polydispersity index 55.5%
According to another exemplary embodiment of the invention, FIGs. 20A - 20D refer to comparative effects of various green-synthesized nanoparticles and extracts on % membrane stabilization and hemolysis inhibition.
FIG. 20A refers to a graph 2000 depicting effects of the AuNPs, the PdNPs, and Diclofenac sodium on % membrane hemolysis inhibition. FIG. 20B refers to a graph 2002 depicting effects of Hexane extract of SCF, Ethanolic extract of SCF, and Diclofenac sodium on % membrane hemolysis inhibition.
FIG. 20C refers to a graph 2004 depicting effects of the AuNPs, the PdNPs, and Diclofenac sodium on % membrane stabilization of HRBC. FIG. 20D refers to a graph 2006 depicting effects of Hexane extract of SCF, Ethanolic extract of SCF, and Diclofenac sodium on % membrane stabilization of HRBC.
Table 3:
S.NO Diclofenac Sodium µg/ml AuNPs PdNPs
Concentration (µg/ml) Mean SD Mean SD Mean SD
1 0.5 80.2 ±0.216 89.5 ±0.124 88.0 ±0.205
2 10 74.2 ±0.249 77.1 ±0.169 81.2 ±0.124
3 20 69.3 ±0.124 68.4 ±0.081 72.2 ±0.205
4 40 52.4 ±0.205 57.4 ±0.262 63.1 ±0.216
5 60 44.3 ±0.169 48.3 ±0.216 52.4 ±0.081
6 80 37.3 ±0.094 39.2 ±0.368 42.3 ±0.124
7 100 21.0 ±0.141 31.0 ±0.216 24.4 ±0.124
In one embodiment, Table 4 hemolytic analysis of hexane and ethanolic extracts diclofenac sodium on hemolysis of Hrbc.
Table 4:
S.NO SD mg/ml Hexane Ethanol
Concentration of SD (mg/ml) Mean SD Mean SD Mean SD
1 0.5 79.07 ±0.205 82.4 ±0.163 79.2 ±0.496
2 10 66.43 ±0.124 75.2 ±0.205 67.1 ±0.339
3 20 58.27 ±0.339 63.0 ±0.169 53.1 ±0.081
4 40 47.23 ±0.286 54.1 ±0.169 49.2 ±0.205
5 60 37.27 ±0.249 42.2 ±0.249 35.0 ±0.492
6 80 24.10 ±0.081 35.5 ±0.081 27.2 ±0.249
7 100 11.60 ±0.864 21.3 ±0.286 16.2 ±0.216
In one embodiment, Table 5 effect of green synthesised nanoparticles AuNPs, PdNPs, and diclofenac sodium on % membrane stabilization.
Table 5:
S.NO Diclofenac Sodium µg/ml AuNPs PdNPs
Concentration (µg/ml) Mean SD Mean SD Mean SD
1 0.5 19.8 ±0.264 10.53 ±0.152 11.03 ±0.251
2 10 25.8 ±0.305 22.93 ±0.208 18.8 ±0.152
3 20 30.7 ±0.152 31.6 ±0.099 27.8 ±0.251
4 40 47.6 ±0.251 42.63 ±0.321 36.9 ±0.264
5 60 55.6 ±0.208 51.7 ±0.264 47.5 ±0.099
6 80 62.6 ±0.115 60.83 ±0.450 57.7 ±0.152
7 100 78.9 ±0.173 72.7 ±0.299 68.7 ±0.264
Standard AuNP PdNP
p-value <0.0001 <0.0001 <0.0001
P value summary **** **** ****
logIC50 2.874 2.177 2.582
t-test t=1.502, df=12 t=0.7910, df=12
Two ANOVA P Value <0.0001 <0.0001 <0.0001
TwowayANOVA P-Value **** **** ****
In one embodiment, Table 6 refers to effect of green synthesised nanoparticles AuNPs, PdNPs, and diclofenac sodium on % membrane stabilization
Table 6:
S.NO Diclofenac Sodium mg/ml Hexane ethanol
Concentration (mg/ml) Mean SD Mean SD Mean SD
1 0.5 20.9 ±0.251 17.6 ±0.200 20.8 ±0.608
2 10 33.6 ±0.152 24.8 ±0.251 32.9 ±0.416
3 20 41.7 ±0.416 37.0 ±0.208 46.9 ±0.100
4 40 52.8 ±0.351 45.9 ±0.152 50.8 ±0.251
5 60 62.7 ±0.305 57.8 ±0.305 65.0 ±0.602
6 80 75.9 ±0.100 64.5 ±0.099 72.8 ±0.305
7 100 88.4 ±1.058 78.7 ±0.351 83.8 ±0.264
Standard Hexane ethanol
p-value <0.0001 <0.0001 <0.0001
P value summary **** **** ***
logIC50 2.57 2.087 2.375
t-test t=1.819, df=12 t=1.382, df=12
Two ANOVA P Value <0.0001 <0.0001 <0.0001
TwowayANOVA P-Value **** **** ****
In one embodiment, the red blood cell (RBC) membrane shares structural and functional similarities with the lysosomal membrane, which plays a crucial role in the anti-inflammatory response. Maintaining lysosomal membrane stability prevents the release of pro-inflammatory enzymes responsible for tissue inflammation and damage, thereby inhibiting the inflammatory process. Non-Steroidal Anti-Inflammatory Drugs (NSAIDs) exhibit anti-inflammatory properties by stabilizing cell membranes and preventing cellular damage.
In one embodiment, hemolysis is induced by hypotonic stress, where RBCs are placed in a hypotonic solution, leading to fluid and electrolyte leakage from the cell membrane. This leakage caused cell shrinkage and, ultimately, haemoglobin oxidation, which resulted in rupturing of the human RBC (HRBC) membrane. The oxidative stress generated during membrane rupture further induced lipid peroxidation, promoting free radical formation. NSAIDs help mitigate this effect by stabilizing the membrane and preventing the efflux of cellular components. The order of membrane stabilization observed is Standard > SCFAuNP > SCFPdNP > Ethanol > Hexane.
According to Table 5, AuNPs exhibits 10.53%–72.7% stabilization, with higher efficacy than diclofenac sodium at 20 mg/ml (31.6% vs. 30.7%). Palladium nanoparticles (PdNPs) showed the least stabilization among the metallic nanoparticles (11.3%–57.7%), performing slightly below the standard drug. The hexane and ethanolic extracts (Table 6) exhibited membrane stabilization within 17.6%–78.7% and 20.8%–83.8%, respectively, at 0.5–100 mg/mL concentrations. However, neither extract displayed significant activity at lower concentrations (0.5–100 mg/ml).
According to Table 3, the hemolysis inhibition percentages for the standard drug (diclofenac sodium) ranged from 80.2% to 21% across the 0.5–100 mg/ml concentration range. Comparatively, the inhibition observed with AuNPs (89.5%–27.3%) and PdNPs (89%–31.3%). The ethanol and hexane extracts exhibited 79.2%–16.2% and 82.4%–21.3% hemolysis inhibition, respectively (Table 4). Their mechanism of action likely involves inhibiting the release of lysosomal enzymes from neutrophils at the site of inflammation. These lysosomal constituents, including bactericidal enzymes and proteases, can exacerbate inflammation and tissue damage when released extracellularly.
This study is the first to report the anti-inflammatory potential of green-synthesized SCF AuNPs, and PdNPs, highlighting their potential as natural anti-inflammatory agents. In one embodiment, the protein denaturation assay is performed using a 3 mL reaction mixture, consisting of 0.450 mL of freshly prepared 5% bovine serum albumin (BSA) solution and 0.05 mL of the test samples. The pH of each sample is adjusted to 6.3 using a small amount of 1N hydrochloric acid. The samples are then incubated at 37°C for 20 minutes in an incubator, followed by heating at 57°C for 30 minutes in a water bath. After heating, the samples are cooled under running tap water, and 2.5 mL of phosphate buffer (pH 7.4) is added to each test tube. The turbidity of the reaction mixture is measured spectrophotometrically at 600 nm. A blank/control is prepared using distilled water, without the addition of test samples.
Inflammation is the primary physiological defense mechanism triggered by external stimuli, and protein denaturation plays a key role in initiating inflammation and arthritic diseases. Protein denaturation refers to the loss of tertiary and secondary structures due to external factors such as strong acids or bases, concentrated inorganic salts, organic solvents, or heat. Most biological proteins lose their functional activity upon denaturation, which contributes to inflammatory conditions, including rheumatoid arthritis, diabetes, and cancer. The inflammatory response is driven by cellular activity and the release of lysosomal enzymes in response to external stimuli. The inhibition of protein denaturation can serve as a potential anti-inflammatory strategy. In one embodiment, BSA (bovine serum albumin) is used as the reagent, as it accounts for nearly 60% of all proteins in animal serum. Upon heating, BSA undergoes denaturation, leading to the expression of antigenic determinants associated with Type III hypersensitivity reactions, which contribute to the inflammatory process.
According to another exemplary embodiment of the invention, FIGs. 21A – 21B refer to effect of green-synthesized nanoparticles and extracts derived from SCF on the inhibition of protein denaturation, which serves as an in vitro indicator of anti-inflammatory activity.
FIG. 21A refers to a graph 2100 depicting effects of the AuNPs, the PdNPs, and Diclofenac sodium on the inhibition of protein denaturation. FIG. 21B refers to a graph 2102 depicting effects of Hexane extract of SCF, Ethanolic extract of SCF, and Diclofenac sodium on on the inhibition of protein denaturation.
In one embodiment, Table 3 presents hemolytic analysis of green synthesised nanoparticles AuNPs, PdNPs, and diclofenac sodium on hemolysis of Hrbc at varying concentrations (0.5–100 mg/ml).
In one embodiment, table 7 refers to effect of green synthesised nanoparticles AuNPs, PdNPs, and diclofenac sodium on protein denaturation.
Table 7:
S.NO Diclofenac Sodium µg/ml AuNPs PdNPs
Concentration (µg/ml) Mean SD Mean SD Mean SD
1 0.5 18.60 ±0.30 17.7 ±0.15 15.43 ±0.15
2 10 39.63 ±0.21 23.6 ±0.10 20.63 ±0.21
3 20 48.37 ±0.35 30.3 ±0.32 30.70 ±0.26
4 40 59.37 ±0.25 38.4 ±0.15 37.60 ±0.20
5 60 67.50 ±0.26 46.6 ±0.10 45.37 ±0.25
6 80 79.33 ±0.32 54.2 ±0.21 52.50 ±0.10
7 100 89.53 ±0.38 61.4 ±0.40 59.57 ±0.15
Standard AuNP PdNP
p-value <0.0003 <0.0001 <0.0001
P value summary *** **** ****
Ic50 67.88 247.5 151.1
log dose 1.832 2.394 2.179
t-test t=0.3463, df=12 t=0.1396, df=12
Two ANOVA P Value <0.0001 <0.0001 <0.0001
In one embodiment, from table 7, the absorbance values of the samples increased proportionally with the increase in sample concentrations. The inhibition values of 18.6% (standard), 17.7% (SCF AuNPs), and 15.43% (SCF PdNPs), respectively.
At higher concentrations (10–100 mg/ml), bimetallic nanoparticles exhibited the highest inhibition of 67.5%, outperforming SCF AuNPs (59.57%) and SCF PdNPs (61.4%) at 100 mg/ml. Additionally, bimetallic nanoparticles demonstrated superior inhibition compared to the fruit extracts (hexane and ethanol). The hexane extract showed an inhibition range of 46.9%–87.4%, while the ethanolic extract exhibited 57.4%–87.4% inhibition across the 0.5–100 mg/mL concentration range. The IC50 values (concentration at which 50% inhibition of protein denaturation occurs) are determined using the linear regression equation with an R² value of 0.9, plotted as concentration vs. percentage of denaturation. The IC50 values are as follows, Diclofenac sodium (standard): 67.88 mg/ml (lowest IC50), PdNPs: 151 mg/ml, AuNPs: 247 mg/ml, Ethanol extract: 28.6 mg/mL, and Hexane extract: 254 mg/mL.
In one embodiment, table 8 refers to effect of hexane and ethanolic extracts of SCF and diclofenac sodium on protein denaturation.
Table 8:
S.NO Diclofenac Sodium mg/ml Hexane Ethanol
Concentration (mg/ml) Mean SD Mean SD Mean SD
1 0.5 13.57 ±1.53 46.97 ±2.27 57.47 ±0.49
2 10 16.90 ±1.00 49.03 ±0.15 64.63 ±0.15
3 20 24.57 ±3.79 59.50 ±0.44 69.43 ±0.35
4 40 33.90 ±2.65 63.50 ±0.36 75.90 ±3.61
5 60 43.90 ±3.61 75.97 ±2.66 76.43 ±0.35
6 80 55.57 ±2.08 80.20 ±0.20 79.23 ±0.50
7 100 65.57 ±3.06 87.47 ±0.31 81.43 ±0.55
Standard Hexane ethanol
p-value <0.0005 <0.0001 <0.0028
P value summary *** **** **
Ic50 67.88 254.6 28.96
log dose 1.832 2.406 1.462
t-test t=0.06519, df=12 t=0.6622, df=12
Two ANOVA P Value <0.0001 <0.0001 <0.0001
Inflammation is recognized as the primary physiological defense mechanism in response to various stressors, and protein denaturation is a key underlying factor contributing to both inflammatory and arthritic diseases. Most biological proteins lose their functional activity upon denaturation, which involves the disruption of their secondary and tertiary structures due to external stimuli such as elevated temperatures, concentrated inorganic salts, strong acids or bases, and certain organic solvents. This denaturation process initiates inflammatory cascades, often associated with chronic conditions such as diabetes, rheumatoid arthritis, and cancer. Inhibition of protein denaturation is therefore a viable strategy for mitigating inflammation. The inflammatory response is further propagated through cellular activation, which is often mediated by the release of lysosomal enzymes in response to external insults.
As shown in Table 7, the absorbance values increased proportionally with rising concentrations of the tested samples, indicating concentration-dependent inhibition of protein denaturation. Monometallic gold (Au) and palladium (Pd) nanoparticles exhibited inhibition rates of 17.7% and 15.43%, respectively. In contrast, biogenic nanoparticles synthesized from Syzygium cumini fruit extracts—specifically SCFAuNPs and SCFPdNPs—demonstrated significantly higher inhibition of protein denaturation, achieving 59.57% and 61.4% inhibition at 100 µg/mL, respectively, over a test range of 10–100 µg/mL. Additionally, hexane and ethanolic extracts of the fruit exhibited inhibition in the ranges of 46.9%–87.4% and 57.4%–87.4%, respectively, across a concentration range of 0.5–100 mg/mL.
The half maximal inhibitory concentration (IC₅₀), defined as the concentration required to inhibit 50% of protein denaturation, was determined using linear regression analysis of concentration versus percentage inhibition plots, with correlation coefficients (R²) of approximately 0.9. The standard anti-inflammatory drug, diclofenac sodium, exhibited the lowest IC₅₀ value at 67.88 µg/mL, followed by SCFPdNPs at 151 µg/mL, SCFAuNPs at 247 µg/mL, ethanol extract at 28.6 mg/mL, and hexane extract at 254 mg/mL. Comparatively, diclofenac sodium was found to be approximately 2.5 times more potent than SCFAuNPs, 4.1 times more potent than SCFPdNPs, 2.4 times more potent than the hexane extract, and 2.1 times more potent than the ethanolic extract in terms of protein denaturation inhibition.
In one embodiment, a mixture comprising 2 mL of aqueous methylene blue solution (50 mM) is treated with 0.5 mL of freshly chilled aqueous NaBH₄ solution (5 mM). To this, nanoparticles selected from SCF AuNPs (25 mg/ml), and SCF PdNPs (12.5 mg/ml) are added. The reaction progress is monitored spectroscopically by measuring absorbance at 663 nm over a period of 3-minute intervals. PdNPs achieved 98% degradation in 18 min with a rate constant of 2.256 × 10⁻¹ min⁻¹.
In one embodiment, a solution of methylene blue (50 mM) is exposed to natural sunlight in the presence of AuNPs, and PdNPs at varying concentrations (25, 50, and 100 mg/ml). The percentage of dye degradation is calculated by spectroscopic absorbance reduction at defined time intervals (0–72 hours). SCF AuNPs demonstrated 99% degradation at 100 mg/ml with a rate constant of 8.378 × 10⁻¹ h⁻¹.
In one embodiment, 0.5 mL of nitrobenzene is added to a round-bottomed flask containing 1.5 mL of ethanol, 100 mg of NaBH₄, and one of the following: 1 mg SCF AuNPs, 1.5 mg PdNPs. The reaction mixture is stirred at a room temperature for a time period of at least 30 minutes. Post-reaction, the mixture is extracted and analyzed via TLC and UV- Visible spectroscopy. PdNPs facilitated effective conversion to aniline, while SCF AuNPs showed negligible catalytic reduction.
The catalytic reduction of nitrobenzene is evaluated using green-synthesized nanoparticles comprising SCF AuNPs, and PdNPs. The reaction outcomes are analyzed using thin-layer chromatography (TLC) and UV-Visible spectroscopy. Among the tested nanoparticles, PdNPs demonstrated effective catalytic reduction of nitrobenzene to aniline, whereas SCF AuNPs showed negligible activity, indicating limited or no catalytic reduction capability. These results establish PdNPs as efficient catalysts for nitro compound reduction.
The underlying reaction mechanism is consistent with previously reported hydride transfer pathways, wherein sodium borohydride (NaBH₄) transfers a hydride ion to the nitro group of nitrobenzene to form a nitroso intermediate. This intermediate undergoes further reduction via successive hydride transfers, ultimately yielding the corresponding amine.
In prior studies, palladium nanoparticles (PdNPs) employed at 5 mg (1 mol%) achieved nitrobenzene reduction over a reaction duration of two hours under mild conditions with NaBH₄ (1.5–4 equivalents), yielding derivatives such as 4-methyl-2-nitroaniline and 4-methoxy-2-nitroaniline. In contrast, the present method using PdNPs (1.5 mg) in the presence of NaBH₄ (300 mg) resulted in the complete reduction of nitrobenzene within 30 minutes at room temperature, indicating significantly enhanced catalytic efficiency. The use of PdNPs for the catalytic reduction of nitrobenzene under the described conditions is reported herein for the first time.
FIGs. 21A – 21B, and tables 7 – 8, display the effect of green-synthesized nanoparticles and SCF extracts on protein denaturation inhibition, a standard in vitro marker for anti-inflammatory activity.
According to another exemplary embodiment of the invention, FIGs. 22A – 22B refer to UV- Visible spectra (2200, 2202) showing reduction in peak intensity for methylene blue in the presence of NaBH₄ and PdNPs, and SCF AuNPs (25 mg/ml) over time. PdNPs exhibited the highest degradation efficiency of (98%) with a rate constant of 2.256 × 10⁻¹ min⁻¹ within 18 mins at a concentration of 12.5 μg/mL. With a rate concentration of 25 μg/mL AuNPs (FIG. 22B) showed 95% degradation with a rate constant of 1.373 × 10⁻¹ min⁻¹ in 21 mins. In addition, the smaller size of SCF PdNPs compared to SCF AuNPs (11.9 nm and 13.17 nm, respectively) may enhance scattering efficiency and catalytic activity. Previous studies on MB degradation using AuNPs synthesized from plant extracts (Sesbania grandiflora and Solieria tenuis) reported that the efficiencies of 80%–97% at 1–9 mg concentrations.
PdNPs from lemon peel and Punica granatum peel extract showed efficiencies of 60%–81%. In contrast, the present study is demonstrated that the highest is observed for PdNPs, and AuNP superior catalytic activity at significantly lower concentrations (12.5–25 μg/mL), highlighting their potential for efficient pollutant degradation in environmental applications.
According to another exemplary embodiment of the invention, FIGs. 22C-22D refer to corresponding pseudo-first-order kinetic plots (2204, 2206) indicating the reaction rate constants for each catalyst. PdNPs achieved the highest rate constant of 2.256 × 10⁻¹ min⁻¹.
Referring to FIG. 22C, the graph 2204 represents a first-order kinetic analysis where the natural logarithm of the ratio of absorbance at time t to the initial absorbance. ln(At/A0) is plotted against time. The linear fit of the data, with an equation y=0.2564x+0.555 and R^2 value of 0.8955, indicates that the reaction follows first-order kinetics with respect to the reactant. The negative slope of the line corresponds to the rate constant k=0.2564 min-1. Additionally, the graph mentions a rate constant value of k= 2.256×101 m-1, which may refer to a pseudo-first-order rate constant involving concentration units, suggesting it was derived from experimental conditions involving a constant excess of one reactant. The reasonably high R^2 value indicates a good fit between the experimental data and the first-order kinetic model.
Referring to FIG. 22D, the graph 2206 graph illustrates a first-order kinetic study where the natural logarithm of the ratio of absorbance at time t to the initial absorbance ln(At/A0) is plotted against time in minutes. The linear regression line, given by the equation y=0.1373x+0.3935, has a high correlation coefficient R^2 value of 0.9422, indicating a strong linear relationship and confirming that the reaction follows first-order kinetics. The slope of the line corresponds to the rate constant k=0.1373 min−1. Additionally, the graph reports a rate constant K=1.29×10−1 m-1, which may denote a pseudo-first-order rate constant derived under conditions where one reactant is in large excess. The strong correlation and linearity confirm the suitability of the first-order kinetic model for this reaction system.
According to another exemplary embodiment of the invention, FIGs. 23A – 23C refer to UV-Visible spectra (2300, 2302, 2304) showing progressive degradation of methylene blue under solar irradiation at different time intervals (0 to 72 hours). FIG. 23A refers to a UV-Visible spectra 2300 of AuNPs at a concentration of 25 μg. FIG. 23B refers to a UV-Visible spectra 2302 of AuNPs at a concentration of 50 μg. FIG. 23C refers to a UV-Visible spectra 2304 of AuNPs at a concentration of 100 μg.
According to another exemplary embodiment of the invention, FIGs. 23D – 23F refer to kinetic plots (2306, 2308, 2310) displaying the corresponding pseudo-first-order reaction rate constants. AuNPs at 100 mg/ml exhibited the highest rate constant of 8.378 × 10⁻¹ h⁻¹.
Referring to FIGs. 23A – 23C, AuNPs exhibited degradation efficiencies of 55%, 68%, and 99% for their corresponding concentrations of 25 μg, 50 μg, 100 μg following pseudo-first-order kinetics with rate constants of 1.2796 × 10⁻¹ h⁻¹, 1.66796 × 10⁻¹ h⁻¹, and 8.37796 × 10⁻¹ h⁻¹ (FIGs. 23D – 23F). The presence of localized surface plasmon resonance (LSPR) in AuNPs enhanced light absorption and electron transfer, accelerating the degradation process. Comparative studies using AuNPs synthesized from Pogostemon benghalensis and Lagerstroemia speciosa showed pseudo-first-order rate constants of 0.1758 min⁻¹ and 19.8 × 10⁻² min⁻¹, achieving ≥90% degradation within 8 minutes upon the addition of sodium borohydride whereas SCF AuNps are degraded with efficiency of 99% for 72 h without any addition of reducing agents.
According to another exemplary embodiment of the invention, FIGs. 24A – 24C refer to UV- Visible spectra (2400, 2402, 2404) depicting dye degradation over 72 hours. According to another exemplary embodiment of the invention, FIGs. 24D – 24F refer to kinetic plots (2406, 2408, 2410) showing pseudo-first-order rate constants. PdNPs at 100 mg/ml reached a maximum degradation efficiency of 82%.
PdNPs (FIGs. 24A – 24C) demonstrated degradation efficiencies of 65%, 72%, and 82%, with rate constants of 1.923 × 10⁻¹ h⁻¹, 3.869 × 10⁻¹ h⁻¹, and 3.96 × 10⁻¹ h⁻¹( FIGs. 24D – 24F), following a dose-dependent trend. Previous studies using SCF PdNPs synthesized from Sapium sebiferum reported 90% of MB degradation within 70 minutes using 9 mg of PdNPs7. In this study PdNPs demonstrated 82% of significant degradation at a concentration of 100μg.
Numerous advantages of the present disclosure may be apparent from the discussion above. In accordance with the present disclosure, a development of plant-mediated gold nanoparticles (AuNPs), and palladium nanoparticles (PdNPs) is disclosed for anti-inflammatory and therapeutic uses.
Biogenic gold nanoparticles (AuNPs), and palladium nanoparticles (PdNPs), have gained significant attention in nanotechnology due to their lower toxicity and eco-friendliness compared to chemically synthesized nanoparticles. These nanoparticles can be synthesized using biological sources such as plants, fungi, and bacteria. In one embodiment, Sarcolobus carinatus aqueous extract is utilized for the biosynthesis of AuNPs, and PdNPs, which exhibited SPR peaks at 545 nm indicating successful nanoparticle formation. These nanoparticles demonstrated good stability at room temperature without additional energy sources. Stability tests showed that SCF AuNPs remained stable for 194 days at room temperature, 120 days at 4°C, and 9 days at -20°C. SCF PdNPs exhibited stability for 135 days at room temperature, 105 days at 4°C, and 11 days at -20°C.
SCF Zeta potential measurements indicated surface charges of -19.4 mV (SCF AuNPs), and -12.6 mV (SCF PdNPs), suggesting moderate stability. The polydispersity index (PDI) values are 33.2% (SCF AuNPs), 55.5% and (SCF PdNPs), while their hydrodynamic sizes are 287 nm, 619.9 nm, and 336 nm, respectively, indicating moderate dispersion. Structural analysis through SEM and EDS confirmed the morphology and elemental composition of the nanoparticles. HRTEM analysis, along with ImageJ software, determined the nanoparticle sizes as AuNPs: 13.17 nm, and PdNPs: 11.9 nm.
Furthermore, XRD spectral analysis confirmed the crystallite structures of these nanoparticles. The anti-inflammatory activity of the biosynthesized AuNPs, and PdNPs along with hexane and ethanolic fruit extracts of Sarcolobus carinatus, is evaluated using the protein denaturation assay and HRBC membrane stabilization assay, with Diclofenac sodium as the reference standard. In the protein denaturation assay, the IC50 values are AuNPs: 247 mg/ml, PdNPs: 151 mg/ml, Standard Diclofenac sodium: 67.88 mg/ml, Hexane extract: 254 mg/mL, and Ethanolic extract: 289 mg/mL. In the membrane stabilization assay, the standard Diclofenac sodium (25.8%), SCF AuNPs (22.9%), and SCF PdNPs (18.8%).
The biosynthesis of AuNPs and PdNPs using Sarcolobus carinatus eliminates the need for toxic chemical reagents, making the process environmentally sustainable. The presence of bioactive compounds such as alkaloids, flavonoids, tannins, and saponins enhances the stability and potential biomedical applications of the nanoparticles.
The synthesized nanoparticles are characterized by controlled shapes (spherical, triangular, hexagonal) and small average particle sizes (AuNPs: 13.17 nm, PdNPs: 11.9 nm), enhancing their applicability in nano-medicine. XRD and Zeta potential analyses confirm the crystalline structure and stability of the nanoparticles, with surface charge values of -19.3 mV (AuNPs), and -10 mV (PdNPs), indicating good colloidal stability.
It will readily be apparent that numerous modifications and alterations can be made to the processes described in the foregoing examples without departing from the principles underlying the invention, and all such modifications and alterations are intended to be embraced by this application.
, Claims:CLAIMS:
I / We Claim:
1. A method for synthesizing nanoparticles (NPs), comprising:
boiling a fruit paste in water using a bain- marie method, followed by filtration for preparing an aqueous fruit extract of Sarcolobus carinatus;
mixing the aqueous fruit extract with an aqueous solution at a ratio of at least 1:6 to obtain a nanoparticles mixture, wherein the aqueous solution comprises at least one of chloroauric acid (HAuCl₄.3H₂O), and palladium chloride (PdCl₂); and
incubating the nanoparticles mixture at a room temperature for a time period to formulate the nanoparticles (NPs),
wherein the nanoparticles (NPs) comprise at least one of gold nanoparticles (AuNPs), and palladium nanoparticles (PdNPs) based on the aqueous solution used, whereby the nanoparticles (NPs) are formulated as an injectable, topical, or oral pharmaceutical composition for inflammatory conditions.
2. The method as claimed in claim 1, wherein the nanoparticles mixture of the aqueous fruit extract with the aqueous solution of chloroauric acid (HAuCl₄.3H₂O) is incubated at the room temperature for a time period of at least 5 hr to obtain the AuNPs.
3. The method for synthesizing nanoparticles (NPs) as claimed in claim 1, wherein the synthesized AuNPs exhibit a surface plasmon resonance (SPR) peak at 545 nm.
4. The method for synthesizing nanoparticles (NPs) as claimed in claim 1, wherein the AuNPs are synthesized at pH values varies between 4 and 12 and remain stable for at least 37 days at a room temperature.
5. The method for synthesizing nanoparticles (NPs) as claimed in claim 1, wherein the AuNPs are separated by centrifugation at 10,000 rpm for a time period of at least 15 min and subsequently purified to obtain pellet.
6. The method for synthesizing nanoparticles (NPs) as claimed in claim 1, wherein the nanoparticles mixture of the aqueous fruit extract with the aqueous solution of palladium chloride (PdCl₂) is incubated at a room temperature for a time period of at least 1 hr to obtain the PdNPs.
| # | Name | Date |
|---|---|---|
| 1 | 202541086886-STATEMENT OF UNDERTAKING (FORM 3) [12-09-2025(online)].pdf | 2025-09-12 |
| 2 | 202541086886-REQUEST FOR EXAMINATION (FORM-18) [12-09-2025(online)].pdf | 2025-09-12 |
| 3 | 202541086886-REQUEST FOR EARLY PUBLICATION(FORM-9) [12-09-2025(online)].pdf | 2025-09-12 |
| 4 | 202541086886-POWER OF AUTHORITY [12-09-2025(online)].pdf | 2025-09-12 |
| 5 | 202541086886-FORM-9 [12-09-2025(online)].pdf | 2025-09-12 |
| 6 | 202541086886-FORM FOR SMALL ENTITY(FORM-28) [12-09-2025(online)].pdf | 2025-09-12 |
| 7 | 202541086886-FORM 18 [12-09-2025(online)].pdf | 2025-09-12 |
| 8 | 202541086886-FORM 1 [12-09-2025(online)].pdf | 2025-09-12 |
| 9 | 202541086886-EVIDENCE FOR REGISTRATION UNDER SSI(FORM-28) [12-09-2025(online)].pdf | 2025-09-12 |
| 10 | 202541086886-EVIDENCE FOR REGISTRATION UNDER SSI [12-09-2025(online)].pdf | 2025-09-12 |
| 11 | 202541086886-EDUCATIONAL INSTITUTION(S) [12-09-2025(online)].pdf | 2025-09-12 |
| 12 | 202541086886-DRAWINGS [12-09-2025(online)].pdf | 2025-09-12 |
| 13 | 202541086886-DECLARATION OF INVENTORSHIP (FORM 5) [12-09-2025(online)].pdf | 2025-09-12 |
| 14 | 202541086886-COMPLETE SPECIFICATION [12-09-2025(online)].pdf | 2025-09-12 |