Skip to content
BY 4.0 license Open Access Published by De Gruyter March 5, 2024

Ascorbic acid-mediated selenium nanoparticles as potential antihyperuricemic, antioxidant, anticoagulant, and thrombolytic agents

  • Muhammad Aamir Ramzan Siddique , Muhammad Aslam Khan , Syed Ali Imran Bokhari EMAIL logo , Muhammad Ismail EMAIL logo , Khurshid Ahmad , Hafiz Abdul Haseeb , Mustafa Mubin Kayani , Suleman Khan , Nafeesa Zahid and Sher Bahadar Khan

Abstract

Selenium (Se) is an important trace element that is involved in controlling oxidative stress and inflammatory disorders. Gouty arthritis is the inflammation and pain within the joints and tissues caused due to the accumulation of monosodium urate (MSU) crystals. This study aimed to investigate the antigout, antioxidant, anticoagulant, and thrombolytic potential of ascorbic acid-mediated Se nanoparticles (A-SeNPs). Different analytical techniques were used to investigate the formation of A-SeNPs. The antigout potential of the nanoparticles was carried out using MSU crystal dissolution, uric acid (UA) degradation assay, and xanthine oxidase inhibition (XOI). A-SeNPs exhibited excellent antihyperurecemic activity in a concentration-dependent manner. It was observed that at the tested concentration of 20 mg·mL−1, the A-SeNPs demonstrated significant breakage and dissolution of MSU crystals and resulted in UA degradation of 67.76%. Similarly, A-SeNPs resulted in 76% XOI with an excellent IC50 of 140 µg·mL−1. Furthermore, considerable antioxidant activity was noted for the A-SeNPs as evaluated with multiple antioxidant assays. Finally, the NPs were found to have significant anticoagulant and thrombolytic potential. Thus, it was concluded that A-SeNPs have potent antihyperuricemic, antioxidant, anticoagulant, and thrombolytic activities, making them an ideal choice for future biomedical applications.

Graphical abstract

1 Introduction

Nanotechnology has attracted the most attention in recent years due to the development of innovative types of materials with distinct physical and chemical properties that are efficiently used in many different areas of research [1,2]. Because of their greater surface-to-volume ratio and higher surface energy, nano-sized particles have distinct characteristics [3,4]. Selenium is an essential element that the human body needs and has shown promising results in the treatment and diagnosis of numerous illnesses [5]. Selenium is a micronutrient in plants, animals, and humans and plays key biological functions [6]. Selenium is a component of proteins and enzymes, such as glutathione peroxidase (GPx), selenoprotein N, thyroxine 5-deiodinase, selenoprotein, and P selenoprotein K [7]. The trace element is required for the biosynthesis of the amino acid selenocysteine and acts as a key cofactor in selenoenzymes that protects the human body against free radical species. Lower Se levels in the human body have been associated with weak immunity, cognitive decline, and high risks of mortality. Recently, there has been growing interest in the preparation and study of SeNPs, because of their excellent mechanical characteristics, optoelectronic and magnetic properties, as well as a plethora of applications in nano-medicines, sensors, and catalysis to name a few [8,9]. SeNPs have also been found to exhibit potential biological activities, reduced toxicities, and significant bioavailability. The NPs are found to be involved in the antioxidant defense system of cells and protect against oxidative stress [10,11]. Elemental Se in the form of NPs also exhibits antimicrobial activities and has shown to possess excellent growth inhibition potential against Candida albicans, Staphylococcus aureus, and Pseudomonas aeruginosa, the pathogens that are associated with hospital and medical devices-acquired infections. Some studies even show that SeNPs exhibit reduced toxicities and are more effective as compared to AgNPs [12,13]. Se is a trace mineral that is essential for maintaining human health. Adults need between 40 and 300 mg of Se daily as a dietary supplement, and it has been linked to more than 40 diseases in humans. SeNPs are naturally broken down by the body, and the resulting Se nutritional supply is harmless to humans [14,15]. SeNPs have been employed in the treatment of cancer [16], drug delivery [17], antibacterial agents [18], antiviral [19], antifungal [9], antioxidant [20], and fertilizers [21]. To date, different fabrication approaches such as physical, chemical, and biological methods have been used to prepare SeNPs with varied morphologies and physicochemical characteristics [22]. However, these approaches particularly chemical and physical methods are considered disadvantageous because of high energy consumption, use of hazardous chemicals, and the non-ecofriendly nature of the methods [23,24]. El Saied et al. successfully investigated and synthesized SeNPs using cell-free extract of the microalgae, Spirulina platensis. The innovative SeNPs used were recommended to generate effectual bioactive agents to control hazardous bacterial species [25]. Srivastava and Mukhopadhyay used the non-pathogenic bacterium Zooglea ramigera to biosynthesize SeNPs with sizes ranging from 30 to 150 nm. Se oxyanions are employed in Zooglea ramigera growth media, and bacterial proteins and enzymes are responsible for Se oxyanions reduction [26]. Similarly, the bacterium Pantoea agglomerans strain UC32 was employed for selenite [Se(iv)] bioreduction to manufacture SeNPs smaller than 100 nm [27]. Wadhwani et al. reported the manufacture of SeNPs from Acinetobacter sp. SW30 cell suspensions and complete cell protein by decreasing Na2SeO3 [28]. The bacterium Klebsiella pneumoniae was used to synthesize SeNPs with a size range of 100–550 nm for the reduction of Se chloride (Se2Cl2) [29].

Besides, SeNPs prepared from naturally existing substances such as ascorbic acid (AA) are comparatively less toxic than SeNPs produced from other physical and chemical routes [30]. Basic calcium phosphate, MSU crystal, and calcium pyrophosphate are identified to produce arthropathies. It is developing proof that these crystals are responsible for the spread of synovitis, cartilage injury, and the development of gout and joint damage [31]. Gout is the main example of arthritis related to MSU crystal deposition in joints and is characterized by hyperuricemia. Inflammatory arthritis gout occurs due to the formation of MSU that results from increased UA levels in the blood [32]. Increased UA levels not only cause inflammatory arthritis but have also been linked to risk factors for various other diseases [33]. To treat gout, anti-inflammatory drugs such as colchicine are prescribed as the first-line treatment option [34]. However, such drugs have a little healing frame and may result in noxious effects [35]. Likewise, urate-lowering drugs (xanthine oxidase inhibitors) such as Allopurinol are also suggested to patients; however, prolonged use of the drug has been associated with severe skin reactions, painful or bloody urination, and liver problems. Thus, safe and effective treatment of hyperuricemia is still unsatisfactory and in need of urgent development [36,37].

The current study aims to use a simple and green route for the synthesis of SeNPs for multiple biomedical applications, including antioxidants, antigout, and anticoagulant studies. The A-SeNPs were synthesized by an eco-friendly approach via the reduction of sodium selenite with L-AA. Then, the synthesized NPs were characterized by X-ray diffraction (XRD) analysis, Scanning electron microscopy (SEM), and Fourier transform infrared (FTIR) spectroscopy, and after characterization, the NPs were explored for multiple biological properties.

2 Experimental

2.1 Materials and methods

The chemicals used were L-Ascorbic acid 99.7% and trichloroacetic acid 99.5% (VWR Chemicals, AnalaR Normapur), sodium selenite Na2SeO3 (Sigma Aldrich, Germany), sodium hydroxide 98% (Daejung Chemicals Siheung-si, South Korea), 2,2-diphenyl-1-picrylhydrazyl (Sigma Aldrich, Germany), uric acid 98% (Avonchem Macclecfield, UK), sodium phosphate 98% (Icon Chemicals) and sulfuric acid 96% (Daejung Chemicals Siheung-si, South Korea). All the chemicals and reagents are used as such without further purification.

2.2 Synthesis of SeNPs

For the synthesis of SeNPs, 2 g of L-AA was dissolved in 100 mL of distilled water (dH2O). In a separate flask, a 10 mmol precursor salt solution of sodium selenite (Na2SeO3-Sigma Aldrich) was prepared in 100 mL of dH2O at 70℃. The prepared L-AA solution was added dropwise to the salt solution, with continuous stirring at 70℃. The transparent sodium selenite solution slowly changed to a reddish color, which indicated the successful synthesis of SeNPs. Afterward, the obtained product was collected and washed three times with dH2O via centrifugation at 4,500 rpm for 15 min and oven-dried at 80℃ for 7 h. The well-dried material was then ground into a fine powder using pastor and mortar, labeled as ascorbic acid-mediated Se nanoparticles (A-SeNPs), and utilized further for characterization and bioassays.

2.3 Characterization

The prepared A-SeNPs were subjected to various characterizations that included FTIR spectroscopy, XRD, SEM, and energy-dispersive X-ray spectroscopy (EDX). For the identification of functional groups and to monitor the surface chemistry of SeNPs, FTIR spectroscopy in the wave number range of 4,000–400 cm−1 was carried out. To determine the material and phase structure of A-SeNPs, XRD was employed in the 2θ (10–80°) using Panalytical X-pert pro MPD XRD. Furthermore, to examine the particle morphology and size, an SEM was used with various magnifications. The morphology and elemental compositional were determined by using Hitachi SU6600 SEM and EDX. Before the characterization, the sample was dried and pasted on carbon tape and then sputtered through gold.

2.4 Biological applications

2.4.1 In vitro anti-gout study

The in vitro anti-gout study was carried out via MSU crystals degradation assay, UA degradation test, and XOI potential of the NPs.

2.5 MSU crystals degradation assay

2.5.1 Preparation of MSU crystal

MSU crystals were prepared using a previously reported method with minor alterations. Briefly, 0.001 M of sodium hydroxide (NaOH) solution was prepared and heated at 70°C. When the temperature reached 70°C, 1.68 g UA dissolved in 50 mL was added dropwise to the NaOH solution. The pH of the mixture was maintained at 7.2. When the dissolution was ensured, the mixture was allowed to cool down at room temperature for 24 h. The supernatant was discarded and the suspension was washed with dH2O thrice. After drying the suspensions, needle-shaped MSU crystals were observed under a microscope and were used for further studies [38].

2.5.2 Preparation of positive control

Two solutions were prepared; A1 is the buffer solution containing 50 mmol·L−1 phosphate (pH 7.4) and 4 mmol·L−1 2-4 dichlorophenol sulfonate (DCPS) was prepared. A2 is the enzyme solution, which was prepared that contained 60 µ·L−1 uricase, 660 µ·L−1 peroxidase (POD), 200 µ·L−1 ascorbate oxidase, and 1 mmol·L−1 4-aminophenazone (4-AP). Both the A1 and A2 mixtures were mixed in 1:1 and were used as a positive control.

2.5.3 Preparation of negative control

A 4 mg·mL−1 MSU crystal in dH2O was used as a negative control.

2.5.4 Effects of SeNPs on crystal dissolution

Four different concentrations of 20, 10, 5, and 2.5 mg·mL−1 of SeNPs were taken into four different falcon tubes and the NPs were well sonicated. After that, 0.5 mg MSU crystals were added into each tube and gently mixed to avoid physical damage to the crystals [39]. The test tubes were maintained on a shaker at a slow speed of 37℃. Consequently, small aliquots (drops) were taken from each tube after 2, 12, and 24 h and were put on a glass slide to observe the dissolution of crystals under the inverted microscope at different time intervals. Multiple micrographs were taken at 40×. In the assay, MSU crystals and dH2O solution were used as a negative control. Enzymes working solution was used as a positive control.

2.5.5 UA degradation test

The UA degradation potential of the NPs was also investigated using the SPINREACT kit. For the preparation of working reagents, working solutions R1 and R2 were prepared according to the SPINREACT kit protocol.

2.5.6 Preparation of working reagents

For the UA degradation test, R1 (buffer solution) and R2 (enzyme solution) were used as working reagents. The working reagent R1 constituted 50 mmol·L−1 phosphate buffer with PH 7.4 and 4 mmol·L−1 DCPS and R2 consisted of 60 U·L−1 uricase, 660 U·L−1 POD, 200 U·L−1 ascorbate oxidase, and 1 mmol·L−1 4-AP. According to the protocol, both the working reagents (R1 and R2) were mixed in 1:1 (v/v). A 6 mg·dL−1 UA aqueous primary standard is ready-made and is provided within the kit.

2.5.7 UA degradation assay

The assay was carried out by mixing 1 mLWR reagents, 25 µL MSU crystal solution (available in the kit) with 25 µL A-SeNPs with varied concentrations, i.e., 2.5–20 mg·mL−1 and the test tubes were incubated for 5 min at 37℃ followed by absorbance recording at 520 nm. In the study, a blank was also used that contained 1 mL working reagents R1 and R2 only. A standard containing 1 mL working reagents (WL) and 25 µL UA aqueous primary standard (6 mg·dL−1) was also employed in the assay. The % UA degradation was calculated using the formula:

(1) % UA degradation = ( sample blank ) ( standard blank ) × 6 ( standard concentration )

2.6 XOI assay

The colorimetric XOI assay that is based on the formation of UA from xanthine was employed as the reference method for the XO inhibition potential of A-SeNPs [40]. In brief, 30 µL of the test sample with different concentrations (25–200 µg·mL−1) was mixed with 210 µL of the phosphate buffer in a test tube. Subsequently, 180 µL of the freshly prepared XO was gently added to the reaction mixture followed by an incubation period of 20 min at 25℃. After the brief incubation period, 960 µL of xanthine was added to the reaction mixture, and the mixture was re-incubated at 25℃ for 20 min, followed by absorbance measurement at 293 nm and calculation of % XOI as

(2) % XOI inhibition = ( A B ) ( C D ) ( A B ) × 100 %

where A depicts enzymatic activity without the presence of sample/positive control and B represents Blank, that is the optical density of the reaction mixture without xanthine oxidase (XO) or the test samples/positive control. C is the enzymatic activity in the presence of sample or positive control and D is the optical density of sample/positive control without the presence of XO.

2.7 Antioxidant study

2.7.1 Total antioxidant capacity

The different concentration of NPs was evaluated for total antioxidant capacity (TAC) by using a phosphomolybdenum-based analysis technique according to the previously reported protocol in the literature with some modifications [41]. In brief, 100 µL aliquot of the sample was carefully added in an Eppendorf tube by using a micropipette and mixed with 900 mL of TAC reagent (which collectively contains 0.6 M sulfuric acid, 28 mM sodium phosphate, and 4 mM ammonium molybdate, in 50 mL dH2O. The mixture was followed by water bath incubation for 45 min, at 80°C, and allowed to cool. After that, the mixture was subjected to the measurement of absorption at the wavelength of 630 nm by using a microplate reader. The TAC was expressed as µg AA equivalent per mg of the NP weights (µg AAE·mg−1).

2.7.2 Total reducing power

The different concentrations of NPs were analyzed for total reducing power (TRP) which was carried out by using the ferric reducing analysis technique in a way of well-optimized reported protocol in the literature with some modifications [42]. Briefly, in an Eppendorf tube, 100 µL aliquot of the sample, 400 µL phosphate buffer (pH 6.6), and 100 µL of potassium ferric cyanide (1% w/v) were added and mixed. The reaction mixture was then incubated by using a water bath for 30 min at 55°C. After that, 200 µL of trichloroacetic acid (10% w/v) was added to the reaction mixture and mixed gently. It was followed by centrifugation at 4,000 rpm for 12 min. Then, an aliquot of 140 µL was taken from the supernatant and poured carefully into the corresponding well in the 96-well plates in which 60 µL of the ferric cyanide solution (0.1% w/v) was already added. After that, the reaction mixture was followed by a measurement of absorption at the wavelength of 630 nm by using a microplate reader. The TRP was calculated as µg AAE·mg−1.

2.7.3 Free radical scavenging assay (FRSA)

NPs were evaluated for 2,2-diphenyl-1-picrylhydrazyl (DPPH) free radical scavenging activity which was done by way of a well-optimized protocol that was reported in the literature, by using a standard bioassay in multiple concentrations that range from 12.5 to 400 µL [43]. The free radical scavenging activity was determined based on the discoloration of the purple color of the DPPH solution. Briefly, in the experiment, an aliquot of 10 µL of the sample was added to the wells of 96-well plates, and a DPPH reagent of 190 µL was mixed with it. After that, it was followed by dark incubation for 1 h at 37°C, and absorbance was measured at the wavelength of 525 nm by using a microplate reader. Free radical scavenging activity was measured as percent (%) inhibition which was calculated by using the given equation:

(3) ( % ) FRSA = 1 Abs sample Abs negative control × 100

where Abs indicates absorbance.

The ability of A-SeNPs to scavenge free radicals was also assessed using the ABTS (2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid)) radical cation bleaching assay [44]. Briefly, ABTS + cationic radical solution was prepared by mixing equal amounts of 2.45 mM potassium persulfate and 7 mM ABTS solution (dH2O), (1:1) and then kept on dark incubation 4℃ 12 h before use. The assay was performed on 96-well plates containing 10 µL of the test sample (4 mg·mL−1, DMSO) and 190 µL of ABTS + solution in each well. Abs was measured at 730 nm and % ABTS was calculated using Eq. 3.

2.8 Anticoagulant and thrombolytic study

Anticoagulant activity of the synthesized SeNPs was evaluated using in vitro co-incubation of the sample with freshly isolated human blood from a healthy person. The isolated blood was co-incubated instantly with varied concentrations of NPs (2.5, 5, 10, and 20 mg·mL−1) in plastic vials (1–8 mL) at room temperature. In the experiment, 100 µL of the sample and 900 µL of the fresh blood were gently mixed, followed by visual observation at different time intervals (10 min, 30 min, 2 h, and 6 h) at room temperature for clot formation. The thrombolytic potential of the A-SeNPs was also determined, by observing the clot dissolution property of A-SeNPs. In the experiment setup, 2–3 drops of freshly isolated human blood were poured onto a sterile clean glass slide and allowed the clot formation. When the clot was ensured, 500 µL of A-SeNPs (20 mg·mL−1) was poured onto the clot. The clot was then observed visually for 1 h to gain an insight into the various stages of thrombolysis [45].

3 Results

3.1 Synthesis of SeNPs

AA was used as a reducing and stabilizing agent during the safe synthesis of A-SeNPs. The AA solution was combined with the stock solution of sodium selenite. The reaction mixture was initially bright; later, the color started to alter, and eventually, brick brick-red color developed, confirming the synthesis of A-SeNPs. It is crucial to be able to produce A-SeNPs from AA since doing so requires less processing afterward. The color of Na2SeO3 solution changed from translucent to brick red during the reduction reaction, indicating the formation of NPs [46]. In our work, the color of the Na2SeO3 solution went from colorless to brick red. The aforementioned findings demonstrate that L-AA can be used as both a reducing agent and a stabilizing agent in the reduction of Se2+ to produce well-dispersed A-SeNPs. L-ascorbic acid is strongly polar and highly water soluble. The electrons in the double bond, the hydroxyl group lone pair, and the carbonyl double bond on the lactone ring form a conjugated system, causing it to behave as a vinylogous carboxylic acid [47]. As a result, L-AA’s structural features provide sufficient reducibility for the creation of Se(0) NPs from Se2+ ions. Scheme 1 can be used to express the redox equation involving selenium ions and L-AA.

Scheme 1 
                  The equation for the reduction and formation of A-SeNPs using AA.
Scheme 1

The equation for the reduction and formation of A-SeNPs using AA.

3.2 Characterization

3.2.1 FTIR analysis

A-SeNPs were characterized through FTIR, to examine the chemical bonding on the surface of NPs through several vibrational modes produced during FTIR as shown in Figure 1. The dangling bonds are formed on the surface which makes the NPs chemically more reactive due to the large aspect ratio as compared to the bulk counterpart. To examine the chemical bonding on the surface of NPs, FTIR spectroscopy in the spectral range from 400 to 4,000 cm−1 was used. The spectrum confirmed the absorption peaks at 978, 1,106, 1,375, 1,626, 2,330, 2,360, and 3,386 cm−1.

Figure 1 
                     FTIR analysis of A-SeNPs.
Figure 1

FTIR analysis of A-SeNPs.

3.2.2 Structural analysis

The phase composition and crystal structure of A-SeNPs were determined through XRD. The pattern of the sample (SeNPs) was monocrystalline with a hexagonal shape that was also confirmed through the XRD technique. The average crystallite size of the SeNPs was examined using the Scherrer formula. Figure 2 shows that the sample was examined in the range of 2θ = 10–80°. The intensity of diffraction peaks was observed at 2θ = 23.9, 29.8, 43.7, 45.7, 51.9, and 68.4° with miller indices that correspond to (100), (101), (102), (111), (201) and (211) planes respectively shown [48]. The XRD pattern of SeNPs was confirmed through JCPDS No: 06-0362.

Figure 2 
                     XRD analysis in A-SeNPs.
Figure 2

XRD analysis in A-SeNPs.

3.2.3 Morphology analysis

The SEM displays the morphological conformation of the synthesized A-SeNPs as shown in Figure 3a. From the apparent surface of the SeNPs seem to be approximately hexagonal, circular with agglomerated in shape. The SEM image (Figure 3a) also confirmed the average size of the A-SeNPs, in the range of 30–40 nm. The compositional analysis was confirmed through energy EDX spectra shown in Figure 3b. Such typical spectra of EDX give conformation about the purity of the SeNPs. Similarly, other researchers also studied the morphological results of SeNPs [49].

Figure 3 
                     (a) SEM micrograph of A-SeNPs and (b) EDX spectra of A-SeNPs.
Figure 3

(a) SEM micrograph of A-SeNPs and (b) EDX spectra of A-SeNPs.

Figure 4 
                     (a) Zeta potential measurement of prepared SeNPs and (b) DLS analysis of SeNPs; size: 70–140 nm; average size: 110.3 nm.
Figure 4

(a) Zeta potential measurement of prepared SeNPs and (b) DLS analysis of SeNPs; size: 70–140 nm; average size: 110.3 nm.

3.2.4 Zeta potential and dynamic light scattering (DLS) analysis

It was discovered that the pattern of distribution of the particle size range from 70 to 140 nm, with an average size of 110.3 nm, was obtained using DLS distributing histograms (Figure 4b). The stability of the dispersion of colloidal NPs is assessed using the zeta potential, which measures the ability of NPs to electrostatically repel one another. It controls the interplay between particles in colloidal dispersion. Moreover, the sign value indicates whether the particle’s surface is dominated by positive or negative forces. The monodispersed dispersion of the NPs was found to be more stable, as evidenced by the zeta potential value of −13.9 mV for SeNPs (Figure 4a). The synthesized SeNPs’ zeta potentials revealed that they were mostly negatively charged. Because the reducing agent AA exposes the presence of gross electrostatic forces with the produced SeNPs, the value of the negative charge potential can be determined [50]. There would be a slight tendency for the particles to stick together if every particle in suspension had a positive or negative zeta potential. Otherwise, the particles would tend to repel one another [51]. The great stability of SeNPs in the absence of aggregation formation was likely caused by the negative charge on the particles (Figures 4 and 5).

3.3 Biological applications

3.3.1 Effect of A-SeNPs on MSU crystals

Gout is a collective term, used to represent certain metabolic conditions that arise from the enhanced production and deposition of MSU crystals, in different connective tissues and joints. It is thus made imperative to grow, characterize, and study the effect of therapeutic substances in vitro and in vivo on the MSU crystals [52]. In the study, a concentration-dependent eminent effect was observed on the morphology and desolation of crystals. For instance, A-SeNPs with different concentrations, i.e., 2.5–20 mg·mL−1, were used to investigate the effect of NPs on MSU crystals as a function of incubation time. In general, A-SeNPs resulted not only in breakage, deformity, and reduction of the size of the MSU crystals but also in desolution. For instance, as can be seen in the figure, a significant effect was observed at 20 mg·mL−1 where the A-SeNPs led to the breaking of maximum crystals, rearrangement, and changing the crystals into small, irregular black and transparent dots. The effect was noticed to be more significant after 12 and 24 h. A similar, but comparatively less effective response was also induced in MSU crystals at 10 mg·mL−1. The effect was found to be alleviated when the concentration of the NPs was decreased as can be observed in Figure 5.

Figure 5 
                     Pictorial presentation of micrographs (40×) showing the effect of A-SeNPs on MSU at varying concentrations. PC denotes a positive control and NC represents a negative control.
Figure 5

Pictorial presentation of micrographs (40×) showing the effect of A-SeNPs on MSU at varying concentrations. PC denotes a positive control and NC represents a negative control.

Contrary to the positive control and NPs, the MSU crystals in the negative control setup retained their needle-shaped triclinic structure even after 24 h. Our study thus concludes that A-SeNPs, affect the structural integrity and result in the desolation of MSU crystals in a dose-dependent pattern.

3.3.2 UA degradation activity

UA in the human body is commonly produced as an end product of purine metabolism. Elevated levels of UA are associated with different complications such as urinary stones and gout arthritis [53]. In the current A-SeNPs with various concentrations were investigated for in vitro quantitative degradation of UA using a SPINREACT (Spain) kit. The results are summarized in Figure 6. In general, it was observed that the UA degrading ability of SeNPs is concentration-depended. For instance, at 20 mg·mL−1, A-SeNPs showed maximum UA degradation of 60.76%. However, the % degradation decreased by reducing the NP concentration, and at the lowest tested concentration of 2.5 mg·mL−1, the UA was degraded up to 39.52%.

Figure 6 
                     Percent degradation of UA by A-SeNPs.
Figure 6

Percent degradation of UA by A-SeNPs.

3.3.3 XOI study

XO or xanthine dehydrogenase serves as an important enzyme that is involved in the conversion of xanthine into UA, leading to hyperuricemia and thus resulting in UA deposition in the joints [54]. In the current study, A-SeNPs were investigated against XOI activity with multiple concentrations ranging from 25 to 200 µg·mL−1. In general, a significant but concentration-dependent XO inhibition activity of the NPs was observed as illustrated in Figure 7. For instance, at the highest tested concentration of 200 µg·mL−1, the NPs resulted in 76% XO inhibition activity as compared to Allopurinol which resulted in 93.2% XO inhibition. When the concentration of A-SeNPs was decreased the inhibitory activity of the NPs also decreased, and at a concentration of 25 µg·mL−1 of NP, only 9.7% inhibition activity was calculated. The IC50 as calculated for the A-SeNPs was found to be 64 µg·mL−1.

Figure 7 
                     % XO inhibition activity of A-SeNPs.
Figure 7

% XO inhibition activity of A-SeNPs.

3.3.4 Antioxidant study

Antioxidants are substances that both scavenge and prevent the synthesis of free radicals, which are created during metabolic events in plants and animals. Cellular damage is caused by greater concentrations of reaction intermediates such as superoxide and hydrogen peroxide [55]. SeNPs regulate reactive oxygen species (ROS) and GPx, which helps reduce free radicals and shield cells from damage. Numerous academic works exist that delineate the antioxidant properties of SeNPs [56,57]. Multiple antioxidant assays, i.e., DPPH and ABTS radical scavenging assays (DPPH-FRSA and ABTS-FRSA), TAC, and ferric-reducing antioxidant power (FRAP), were executed to quantify the in vitro antioxidant potential of A-SeNPs as summarized in Figure 8. ABTS-FRSA (2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) and DPPH-FRSA (2,2-diphenyl-1-picrylhydrazyl) depend on neutralization of the stable colored radicals of DPPH˙ + (purple) and ABTS + (green) by the antioxidant sample in question signifying the potential free radical scavenging ability of the sample [58]. SeNPs stabilized with chitosan showed antioxidant action by enhancing and preserving GPx and inhibiting the production of lipofuscin in mice, according to a study by Zhai et al. [59]. Likewise, phytofabricated SeNPs derived from Emblica officinalis fruit extract showed antioxidant efficacy in DPPH and ABTS experiments [46]. When tested for hydroxyl radical and DPPH radical scavenging, gum-Arabic-stabilized SeNPs showed more antioxidant efficacy than gum-Arabic-hydrolyzed alkali [60].

Figure 8 
                     Antioxidant potential of A-SeNPs.
Figure 8

Antioxidant potential of A-SeNPs.

In the present study, the considerable concentration-dependent free radical scavenging potential of the A-SeNPs was confirmed. For instance, at the maximum tested concentration of 400 µg·mL−1, ABTS-FRSA and DPPH-FRSA were noted as 50.58% ± 1.20 (IC50: 380 µg·mL−1) and 33.97% ± 1.01 (IC50 > 400 µg·mL−1) respectively. The free radical scavenging ability of the NPs was further supported by FRAP and phosphomolybdate-based TAC assays. The mechanism of FRAP or TRP is based on the formation of the ferric–ferrous chromatic complex by the antioxidant sample which can be quantified spectrophotometer at 700 nm while the phosphomolybdenum-based assay depends on the spectrophotometric detection of molybdenum(v) which is formed as reduction of molybdenum(vi) in an acidic environment by the test sample [61]. At the tested concentration of 400 µg·mL−1, maximum, FRAP and TAC activities were found to be 29.66 µg AAE·mg−1 of NPs and 31.27 µg AAE·mg−1 of NPs, respectively. However, the antioxidant potential significantly decreased by reducing the concentration of NPs, thus affirming the concentration-dependent antioxidant capacity of the A-SeNPs.

3.3.5 Anticoagulant and thrombolytic properties

Anticoagulant activity of the A-SeNPs was tested by co-incubation of samples with freshly isolated blood and the clots were visualized as a function of storage time as displayed in Figure 9. The blank containing only fresh blood (NC) began to coagulate soon after the initiation of the experiment and the blood sample, finally formed a thick blood clot till 10 min. On the other hand, the A-SeNPs, with different concentrations, showed excellent anticoagulant activity and the blood samples did not undergo significant changes till 3 h as can be visualized in Figures 9a and b.

Figure 9 
                     Anticoagulant (a) and thrombolytic (b) results of A-SeNPs as a function of time.
Figure 9

Anticoagulant (a) and thrombolytic (b) results of A-SeNPs as a function of time.

The best anticoagulant results were visualized at 20 mg·mL−1 where negligible or no clot can be seen. However, as the concentration of the NPs was reduced, the anticoagulant ability of the test samples was slightly reduced as little or considerable clots can be observed in other test samples, especially after 3 h. Furthermore, the positive control containing ethylenediaminetetraacetic acid (EDTA) also resulted in no clot formation till 3 h, thus affirming and supporting the anticoagulant property of A-SeNPs as observed in the experiment. For the determination of potential thrombolytic properties, dH2O dispersion of A-SeNPs (20 mg·mL−1) was added to a pre-formed blood clot on a clean sterile glass slide, and visual observations were made. With time, eminent clot lysis was observed as shown in Figure 9b.

4 Discussion

There has been growing interest in the preparation and study of SeNPs because of their excellent mechanical characteristics, optoelectronic and magnetic properties, as well as their plethora of applications in nano-medicines, sensors, xerography, and catalysis to name a few [62].

Conventionally, chemical and physical methods have been used to prepare nanoscale Se particles with different morphologies and physicochemical characteristics [63]. However, these approaches, in particular chemical and physical methods, are considered disadvantageous because of high energy consumption, use of hazardous chemicals, and the non-ecofriendly nature of the methods. It has also been reported that the SeNPs synthesized from phytochemicals such as AA is a less toxic and environmentally friendly process than the SeNPs prepared from the chemical and physical process [63,64].

Recent studies show that Se at a nano-scale exhibits far better biocompatibility and lower toxicities than inorganic and organic Se compounds. Such benefits of the SeNPs make them potential nanomaterials to be explored as therapeutic and theranostic agents [65]. In the present study, we planned to investigate the A-SeNPs for potential antigout (MSU crystals desolation, UA degradation, XOI, antioxidant, anticoagulant, and thrombolytic properties.

After successful synthesis, the ascorbic-mediated synthesized SeNPs were characterized using XRD, FTIR, SEM, and EDX.

XRD patterns confirmed the diffraction peaks at 2θ = 23.9, 29.8, 43.7, 45.7, 51.9, and 68.4° with miller indices that correspond to (100), (101), (102), (111), (201), and (211) planes respectively (JCPDS No: 06-0362) confirming the hexagonal structure [66]. The size of the prepared synthesized A-SeNPs was calculated by Scherrer’s equation [67] as given;

(4) D ( nm ) = k λ / β 1 / 2 cos θ

where D represents the size of the NP, k is a constant equal to 0.9, λ used as radiation source from CuKα 0.15432 nm, β1/2 is the full width at half maximum, and 2θ is the diffraction angle. The calculated crystallite size of A-SeNPs was 78 nm. Similar XRD patterns and crystallite dimensions of the nanoscale Se were also reported previously in the literature [66,68].

The surface interaction between the A-SeNPs and stabilizing agent was investigated using FTIR analysis. The FTIR spectra of A-SeNPs confirm the variety of bands present in AA (Figure 1). The broadband at 3,550–3,190 cm−1 corresponds to the O–H hydroxyl group of the AA present on the surface of NPs sample (A-SeNPs) [69,70,71,72]. The two sharp peaks at 2,330 and 2,360 cm−1 correspond to the molecules of CO2 that may exist in the air during the characterization of the SeNPs. The bond at 1,626 cm−1 represents the symmetric C═C double bond stretching motion of dehydroascorbic acid [73]. The bond in a region 1,375 cm−1 is attributed to O–H bend or ring deformation [74]. Figure 1 confirms that peaks at 978 and 1,106 cm−1 are attributed to aromatic in-plane C–H bending and secondary OH, respectively.

Similarly, SEM micrographs revealed the hexagonal, spherical, and slightly agglomerated morphology of the prepared A-SeNPs with a mean size of 30–40 nm. Almost similar morphology of SeNPs was also presented in other studies [75,76]. For instance, in a study, stable, uniform, hexagonal phase and spherical SeNPs with an average size of 30–40 nm were prepared using a green and economical approach. Such SeNPs were found to be highly potent against bacterial and fungal strains [76]. Finally, EDX affirmed the elemental Se within A-SeNPs. After physiochemical, morphological, and elemental confirmation, the A-SeNPs were explored in multiple biological studies including anti-gout, antioxidant, and anticoagulant.

Gout is caused by hyperuricemia, a metabolic disorder characterized by elevated UA in the body due to an imbalance of UA excretion and production [77]. Currently, throughout the world, the prevalence of hyperuricemia has continuously increased manifolds [78]. Presently, a limited number of medicines are available, for treatment and most of these are labeled with side effects. For instance, in European countries since 2003, the uricosuric agent benzbromarone was discontinued due to severe hepatotoxicity reports [79]. While XO inhibitor allopurinol is primarily used for the treatment of hyperuricemia, this drug has severe cutaneous side effects [80]. Hence, a lot of previous research has been conducted to develop safe therapeutic agents for the treatment of elevated UA level-related disorders [79].

In this study, we mainly targeted MSU crystals that are formed through the accumulation of UA in the body leading to gout and UA causal agent of hyperuricemia. Interestingly, A-SeNPs showed remarkable crystal dissolution activity and after 24 h inhibited the reformation of MSU crystal. Few considerations such as extremes of pH and high temperature could be argued to influence the dissolution of MSU crystals. However, the pH of A-SeNPs dispersion was noted as 5.56, which is in fact near neutral, not extremely acidic. Similarly, high temperature may facilitate MSU crystal dissolution; however, the current experiment was executed at the physiological body temperature, i.e., 37°C [81], which thus leads to the only possibility that A-SeNPs played a potential role in the crystal dissolution process.

Recently, various enzymes-NPs that mimic natural enzymes have been explored that may be used as an alternative to uricase (UOX) or in combination with uricase to catalyze the degradation of UA. For instance, Jung and Kwon found that UOX-AuNP successfully degraded UA, five times more rapidly than uricase alone [82]. Similarly, in another study, researchers demonstrated that platinum NPs (PtNPs) with sizes 5–55 nm, mimic the natural uricase, and effectively carry out the oxidative degradation of UA. The authors also concluded that the degradation mechanism was pH-independent and was affected little by the experimented size range of the NPs [83]. In our study, the A-SeNPs were also screened for the UOX mimic characteristic to degrade the UA. We found a concentration-dependent degradation ability of A-SeNPs as it was noted to be a maximum of 20 mg·mL−1 and a minimum of 2.5 mg·mL−1. However, a detailed study is needed to be formulated to investigate the exact mechanism involved in the degradation process of UA. The A-SeNPs were also tested for the XOI potential. XO or xanthine dehydrogenase is a key enzyme in the pathogenesis of gout as it catalyzes the conversion of xanthine into UA, leading to hyperuricemia and thus resulting in UA deposition in the joints. The search for novel potential molecules with XO inhibition activity has grown in the recent past, because of the toxic effects that are being associated with approved drugs like allopurinol and febuxostat [84]. It is now a well-established fact that the interaction of proteins with NPs results in the formation of “protein corona” that may in turn affect the physicochemical properties and nature of both the interacting NPs and proteins. This also forms based on NP bio-reactivity with specific protein molecules [85]. In the study, an excellent but concentration-dependent inhibition activity of A-SeNPs was measured at 400 µg·mL−1 resulting in 76% XO inhibition activity, but the activity decreased considerably at the lowest tested concentration of 50 µg·mL−1 to only 9.7%. Because of the ease of production and low cost, metallic NPs may be considered as a potential alternative to conventional enzyme inhibitors, once their biosafe nature is ensured and thus may prove to be useful tools to treat different pathological conditions such as gout.

It is now a fact that different abiotic stresses build up highly toxic ROS in plant and animal cells that in turn denature key proteins, lipids, carbohydrates, and DNA and ultimately result in oxidative stress [86,87]. Several different diseases may thus arise due to the buildup of oxidative stress. Such oxidative stress-related diseases can be prevented by the intake of antioxidants that function to neutralize the effects of ROS [88]. In the presented study, DPPH and ABTS radical scavenging assays (DPPH-FRSA, ABTS-FRSA), TAC, and FRAP were performed to evaluate the antioxidant capacity of A-SeNPs. The considerable, concentration-dependent free radical scavenging ability of the A-SeNPs was detected as maximum ABTS-FRSA and DPPH-FRSA at 400 µg·mL−1 was noticed as 50.58% ± 1.20 (IC50: 380 µg·mL−1) and 33.97% ± 1.0 (IC50: 400 µg·mL−1), respectively.

As blood comes into contact with subendothelial surfaces, it clots rapidly and stays fluid inside the vasculature. Coagulation / fbrinolysis balance prevents thrombosis and bleeding in normal circumstances. Any imbalance promotes coagulation, resulting in thrombosis, platelet aggregation, fibrin formation, and stuck red blood cells in arteries or veins. Several antithrombotic medications are available on the market to treat thrombosis. Antiplatelet drugs reduce platelet activation or aggregation, while anticoagulants prevent fibrin formation; however, fibrinolytic treatments break down fibrin synthesis [89]. It has recently been shown that using naturally produced AgNPs instead of selective active molecules has effective anticoagulant effects and requires less than active molecules [90]. Using a different strategy, red algae-derived bioinspired cobalt NPs have effectively stopped blood clot formation in vitro [91]. Previous studies on metal oxide NP, especially on gold and silver NP, have helped to clarify the likely biochemical mechanism, even though the mechanisms underlying SeNPs’ anticoagulant activity have not been theoretically deciphered. These investigations showed that these NPs prevent prothrombin from becoming thrombin, which is an essential stage in the creation of insoluble fibrin strands and the catalysis of other clotting factors [92,93]. However limited material in the literature is available about the use of metallic NPs as anticoagulant and thrombolytic agents, especially regarding SeNPs. Therefore, we explored the greenly synthesized SeNPs for anticoagulant and thrombolytic potential. The considerable anticoagulant activity of the NPs was visualized which was mainly found to be concentration-dependent as even till 3 h no eminent clot was observed in the blood sample at 20 mg·mL−1. The findings obtained in the current study, are comparable to the thrombolytic behavior of AgNPs, AuNPs, and bimetallic Au-Ag-NPs [82,94,95]. Our investigation thus revealed the effectiveness of A-SeNPs as potential anticoagulant and thrombolytic agents. The exact mechanism of the thrombolytic activity of a metallic nanostructure has not yet been interpreted, but the probable biochemical mechanism can be explained via the thrombolysis process. The A-SeNPs might have resulted in the inhibition of clot-forming enzymes or prevention of prothrombin transformation into thrombin, which in turn catalyzes the formation of insoluble fibrin and other clotting factors [92]. Traditional antithrombotic drugs such as streptokinase come with a couple of drawbacks including low shelf life, neutralization of foreign materials, and the prospect of unnecessary bleeding [96]. So, the potential anticoagulant behavior of A-SeNPs may have some beneficial medical applications to control thrombosis and other related illnesses. Antihyperuricemic, antioxidant, anticoagulant, and thrombolytic activities of A-SeNPs and their comparison with the recent literature is presented in (Table 1).

Table 1

Antihyperuricemic, antioxidant, anticoagulant, and thrombolytic activities of A-SeNPs and their comparison with the recent literature

S. no Materials used Shape and size Applications Ref
1 Bacillus endophyticus-mediated SeNPs Rod-shaped with an average size from 10 to 20 nm Ba-SeNpMo had 67% thrombolytic activities, compared to 82% for the positive control and 4.0% for the negative control [97]
2 CTAB-SeNPs Irregular with an average size of 32.53 nm DPPH radical scavenging activity of 23.72 ± 008% [98]
SDS-SeNPs, Irregular with an average size of 48.04 nm DPPH radical scavenging activity of 25.14 ± 0.01%
Brij-58-SeNPs Spherical with average size of 37.78 nm DPPH radical scavenging activity of 24.46 ± 0.04%
3 Chitosan-stabilized SeNPs Spherical with an average size of 103 nm SeNPs have a higher ABTS scavenging ability that the value could reach up to 87.45 ± 7.63%. [59]
4 Penicillium expansum-mediated SeNPs Spherical with an average size 4–12.7 nm Concentrations above 30 μg·L−1 of SeNPs have antioxidant activity above 50% [99]
5 Crocus caspius-mediated SeNPs Spherical with a size of 79 nm The IC50 of SeNPs has obtained 123.9 ± 2.5 mg·mL−1. The standard compound was Butylated Hydroxyanisole with an IC50 value of 53.96 mg·mL−1 [100]
6 Diospyros Montana-mediated SeNPs Spherical shape with an average size of 4–16 nm SeNPs showed 61.12% of DPPH scavenging activity at the concentration of 200 μg·mL−1 compared to the AA (99.84%) [68]
7 Tinospora cordifolia-mediated SeNPs Spherical shape with a size range of 20–200 nm SeNPs showed DPPH scavenging activity of 78.03% at 100 μg·mL−1 than aqueous extract of T. cordifolia which showed 50.83% scavenging activity [101]
8 Withania somnifera-mediated SeNPs Spherical with average size of 45–90 nm SeNPs were found to possess significant antioxidant activity (IC50 – 14.81 μg·mg−1) [102]
9 Solanum lycopersicum-mediated SeNPs Irregular with an average size of 1,155 nm Solanum lycopersicum-mediated SeNPs have DPPH antioxidant activity with an IC50 value of 20.7398 mg·mL−1 [103]
10 Citric acid-mediated SeNPs Spherical with size from 10.5 to 20 nm Citric acid-mediated SeNPs exhibit higher anticoagulant properties with increased SeNPs dose but still less than heparin which is the standard anticoagulant drug [93]
11 AA-mediated SeNPs Hexagonal with particle size range from 70 to 140 nm, with an average size of 110.3 nm DPPH 50.58 ± 1.20% (IC50 of 380 µg·mL−1 and ABTS 33.97 ± 1.01% (IC50 > 400 µg·mL−1 Present work
At a concentration of 400 µg·mL−1 showed 76% XO inhibition activity and 9.7% at a concentration of 50 µg·mL−1
Prepared A-SeNPs, showed excellent anticoagulant activity and the blood samples did not undergo significant changes till 3 h

5 Conclusion

In the current study, A-SeNPs were synthesized through a facile precipitation technique by using L-AA as a reducing and stabilizing agent. The L-AA played an important role as a reducing agent by controlling the size of particles. The technique used in the study is simple, non-toxic, inexpensive, and environmentally friendly. Characterization techniques such as FTIR, XRD, EDX, and SEM support the structure, size, and crystallinity of selenium NPs. In the current study, A-SeNPs were very effective against MSU crystal dissolution and UA degradation. After a 2–24 h experiment, it was noted that A-SeNPs considerably dissolved and inhibited the regrowth of MSU crystal at concentrations of 20 and 10 mg·mL−1. Further, the synthesized NPs showed potent inhibition against XO. According to these results, it can be concluded that A-SeNPs have the potency to treat gout and hyperuricemia-related diseases. However, further in vivo studies need to be conducted regarding the efficacy of A-SeNPs to treat such conditions in an in vivo environment. Moreover, the A-SeNPs showed considerable antioxidant potential as evaluated by multiple antioxidant assays. Most importantly, A-SeNPs were found to have good anticoagulant and thrombolytic activities. Our study thus extends the knowledge of biological applications of SeNPs and opens up new opportunities in the area of biomedical research.

Acknowledgments

This article is part of the Ph.D. thesis of Mr. Muhammad Aamir Ramzan Siddique. The authors are very thankful to the Department of Biological Sciences, International Islamic University Islamabad, and HEC Pakistan for providing an opportunity for an IRSIP fellowship for important research to the University of Alberta Canada to complete this research work.

  1. Funding information: Authors state no funding involved.

  2. Author contributions: M.A.R. Siddique and S.A.I. Bokhari perceived the idea, carried out experimental work, and wrote the manuscript. M.A. Khan, K. Ahmed, and H.A. Haseeb assisted in experimental work and manuscript preparation, M. M. Kayani and N. Zahid, assisted in experimental work and characterizations of samples, S. Khan, M. Ismail, S. B. Khan, proofread and assisted the nano work, and S. A. I. Bokhari supervised the project. All authors read and approved the final manuscript.

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

  4. Data availability statement: All data generated or analyzed during this study are included in this published article.

References

[1] Ismail M, Khan MI, Khan SA, Qayum M, Khan MA, Anwar Y, et al. Green synthesis of antibacterial bimetallic Ag–Cu nanoparticles for catalytic reduction of persistent organic pollutants. J Mater Sci: Mater Electron. 2018;29:20840–55.10.1007/s10854-018-0227-2Search in Google Scholar

[2] Albukhari SM, Ismail M, Akhtar K, Danish EY. Catalytic reduction of nitrophenols and dyes using silver nanoparticles @ cellulose polymer paper for the resolution of waste water treatment challenges. Colloids Surf A: Physicochem Eng Asp. 2019;577:548–61.10.1016/j.colsurfa.2019.05.058Search in Google Scholar

[3] Rehman AU, Sharafat U, Gul S, Khan MA, Khan SB, Ismail M, et al. Green synthesis of manganese-doped superparamagnetic iron oxide nanoparticles for the effective removal of Pb(ii) from aqueous solutions. Green Process Synth. 2022;11:287–305.10.1515/gps-2022-0030Search in Google Scholar

[4] Khan SB, Ismail M, Bakhsh EM, Asiri AM. Design of simple and efficient metal nanoparticles templated on ZnO-chitosan coated textile cotton towards the catalytic reduction of organic pollutants. J Ind Text. 2022;51:1703S–28S.10.1177/1528083720931481Search in Google Scholar

[5] Xu T, Lai J, Su J, Chen D, Zhao M, Li Y, et al. Inhibition of H3N2 influenza virus induced apoptosis by selenium nanoparticles with chitosan through ROS-mediated signaling pathways. ACS omega. 2023;8:8473–80.10.1021/acsomega.2c07575Search in Google Scholar PubMed PubMed Central

[6] Zohra E, Ikram M, Omar AA, Hussain M, Satti SH, Raja NI, et al. Potential applications of biogenic selenium nanoparticles in alleviating biotic and abiotic stresses in plants: A comprehensive insight on the mechanistic approach and future perspectives. Green Process Synth. 2021;10:456–75.10.1515/gps-2021-0047Search in Google Scholar

[7] Sun P, Wang Z, Yuan N, Lu Q, Sun L, Li Y, et al. Evaluation of nano-selenium biofortification characteristics of alfalfa (Medicago sativa L.). Green Process Synth. 2023;12:20228121.10.1515/gps-2022-8121Search in Google Scholar

[8] Langi B, Shah C, Singh K, Chaskar A, Kumar M, Bajaj PN. Ionic liquid-induced synthesis of selenium nanoparticles. Mater Res Bull. 2010;45:668–71.10.1016/j.materresbull.2010.03.005Search in Google Scholar

[9] Fardsadegh B, Vaghari H, Mohammad-Jafari R, Najian Y, Jafarizadeh-Malmiri H. Biosynthesis, characterization and antimicrobial activities assessment of fabricated selenium nanoparticles using Pelargonium zonale leaf extract. Green Process Synth. 2019;8:191–8.10.1515/gps-2018-0060Search in Google Scholar

[10] Al-Saggaf MS, Tayel AA, Ghobashy MOI, Alotaibi MA, Alghuthaymi MA, Moussa SH. Phytosynthesis of selenium nanoparticles using the costus extract for bactericidal application against foodborne pathogens. Green Process Synth. 2020;9:477–87.10.1515/gps-2020-0038Search in Google Scholar

[11] Hatami R, Javadi A, Jafarizadeh-Malmiri H. Effectiveness of six different methods in green synthesis of selenium nanoparticles using propolis extract: Screening and characterization. Green Process Synth. 2020;9:685–92.10.1515/gps-2020-0065Search in Google Scholar

[12] Wang Q, Webster TJ. Nanostructured selenium for preventing biofilm formation on polycarbonate medical devices. J Biomed Mater Res Part A. 2012;100:3205–10.10.1002/jbm.a.34262Search in Google Scholar PubMed

[13] Tran PA, Webster TJ. Selenium nanoparticles inhibit Staphylococcus aureus growth. Int J Nanomed. 2011;6:1553–8.10.2147/IJN.S21729Search in Google Scholar PubMed PubMed Central

[14] Al Jahdaly BA, Al-Radadi NS, Eldin GMG, Almahri A, Ahmed MK, Shoueir K, et al. Selenium nanoparticles synthesized using an eco-friendly method: dye decolorization from aqueous solutions, cell viability, antioxidant, and antibacterial effectiveness. J Mater Res Technol. 2021;11:85–97.10.1016/j.jmrt.2020.12.098Search in Google Scholar

[15] Shirsat S, Kadam A, Naushad M, Mane RS. Selenium nanostructures: microbial synthesis and applications. RSC Adv. 2015;5:92799–811.10.1039/C5RA17921ASearch in Google Scholar

[16] El-Hefnawy ME, El-Sherbiny MM, Al Harbi M, Tayel AA. Synergistic in vitro anticancer actions of decorated selenium nanoparticles with fucoidan/Reishi extract against colorectal adenocarcinoma cells. Green Process Synth. 2022;11:373–84.10.1515/gps-2022-0035Search in Google Scholar

[17] Jalalian SH, Ramezani M, Abnous K, Taghdisi SM. Targeted co-delivery of epirubicin and NAS-24 aptamer to cancer cells using selenium nanoparticles for enhancing tumor response in vitro and in vivo. Cancer Lett. 2018;416:87–93.10.1016/j.canlet.2017.12.023Search in Google Scholar PubMed

[18] Borna F, Hoda J-F. Aloe vera leaf extract mediated green synthesis of selenium nanoparticles and assessment of their In vitro antimicrobial activity against spoilage fungi and pathogenic bacteria strains. Green Process Synth. 2019;8:(1)399–407.10.1515/gps-2019-0007Search in Google Scholar

[19] Lin Z, Li Y, Xu T, Guo M, Wang C, Zhao M, et al. Inhibition of enterovirus 71 by selenium nanoparticles loaded with siRNA through bax signaling pathways. ACS omega. 2020;5:12495–500.10.1021/acsomega.0c01382Search in Google Scholar PubMed PubMed Central

[20] Ahmad I, Younas Z, Mashwani ZUR, Raja NI, Akram A. Phytomediated selenium nanoparticles improved physio-morphological, antioxidant, and oil bioactive compounds of sesame under induced biotic stress. ACS Omega. 2023;8:3354–66.10.1021/acsomega.2c07084Search in Google Scholar PubMed PubMed Central

[21] Gudkov SV, Shafeev GA, Glinushkin AP, Shkirin AV, Barmina EV, Rakov II, et al. Production and use of selenium nanoparticles as fertilizers. ACS omega. 2020;5:17767–74.10.1021/acsomega.0c02448Search in Google Scholar PubMed PubMed Central

[22] Wen S, Hui Y, Chuang W. Biosynthesis and antioxidation of nano-selenium using lemon juice as a reducing agent. Green Process Synth. 2021;10:178–88.10.1515/gps-2021-0018Search in Google Scholar

[23] Geoffrion LD, Hesabizadeh T, Medina-Cruz D, Kusper M, Taylor P, Vernet-Crua A, et al. Naked selenium nanoparticles for antibacterial and anticancer treatments. ACS Omega. 2020;5:2660–9.10.1021/acsomega.9b03172Search in Google Scholar PubMed PubMed Central

[24] Ismail M, Khan MI, Akhtar K, Seo J, Khan MA, Asiri AM, et al. Phytosynthesis of silver nanoparticles; naked eye cellulose filter paper dual mechanism sensor for mercury ions and ammonia in aqueous solution. J Mater Sci: Mater Electron. 2019;30:7367–83.10.1007/s10854-019-01049-xSearch in Google Scholar

[25] El Saied BEF, Diab AM, Tayel AA, Alghuthaymi MA, Moussa SH. Potent antibacterial action of phycosynthesized selenium nanoparticles using Spirulina platensis extract. Green Process Synth. 2021;10:49–60.10.1515/gps-2021-0005Search in Google Scholar

[26] Srivastava N, Mukhopadhyay M. Biosynthesis and structural characterization of selenium nanoparticles mediated by Zooglea ramigera. Powder Technol. 2013;244:26–9.10.1016/j.powtec.2013.03.050Search in Google Scholar

[27] Torres SK, Campos VL, León CG, Rodríguez-Llamazares SM, Rojas SM, González M, et al. Biosynthesis of selenium nanoparticles by Pantoea agglomerans and their antioxidant activity. J Nanopart Res. 2012;14:1–9.10.1007/s11051-012-1236-3Search in Google Scholar

[28] Wadhwani SA, Gorain M, Banerjee P, Shedbalkar UU, Singh R, Kundu GC, et al. Green synthesis of selenium nanoparticles using acinetobacter sp. SW30: Optimization, characterization and its anticancer activity in breast cancer cells. Int J Nanomed. 2017;12:6841–55.10.2147/IJN.S139212Search in Google Scholar PubMed PubMed Central

[29] Fesharaki PJ, Nazari P, Shakibaie M, Rezaie S, Banoee M, Abdollahi M, et al. Biosynthesis of selenium nanoparticles using Klebsiella pneumoniae and their recovery by a simple sterilization process. Braz J Microbiol. 2010;41:461–6.10.1590/S1517-83822010000200028Search in Google Scholar

[30] Shar A, Lakhan M, Wang J, Ahmed M, Alali K, Ahmed R, et al. Facile synthesis and characterization of selenium nanoparticles by the hydrothermal approach. Dig J Nanomater Biostruct. 2019;14:867–72.Search in Google Scholar

[31] Whelan LC, Morgan MP, McCarthy GM. “Basic calcium phosphate crystals as a unique therapeutic target in osteoarthritis. Front Bioscience-Landmark. 2005;10:530–41.10.2741/1549Search in Google Scholar PubMed

[32] Cleophas M, Joosten L, Stamp LK, Dalbeth N, Woodward OM, Merriman TR. ABCG2 polymorphisms in gout: insights into disease susceptibility and treatment approaches. Pharmacogenomics Pers Med. 2017;10:129–42.10.2147/PGPM.S105854Search in Google Scholar PubMed PubMed Central

[33] Abeles AM. Hyperuricemia, Gout, and cardiovascular disease: An update. Curr Rheumatol Rep. 2015;17:13.10.1007/s11926-015-0495-2Search in Google Scholar PubMed

[34] Cronstein BN, Terkeltaub R. The inflammatory process of gout and its treatment. Arthritis Res Ther. 2006;8:1–7.10.1186/ar1908Search in Google Scholar PubMed PubMed Central

[35] Finkelstein Y, Aks SE, Hutson JR, Juurlink DN, Nguyen P, Dubnov-Raz G, et al. Colchicine poisoning: the dark side of an ancient drug. Clin Toxicol. 2010;48:407–14.10.3109/15563650.2010.495348Search in Google Scholar PubMed

[36] Stern RJ. Reducing life-threatening allopurinol hypersensitivity. JAMA Intern Med. 2015;175:1558–8.10.1001/jamainternmed.2015.3546Search in Google Scholar PubMed

[37] Richette P, Doherty M, Pascual E, Barskova V, Becce F, Castañeda-Sanabria J, et al. 2016 updated EULAR evidence-based recommendations for the management of gout. Ann Rheum Dis. 2017;76:29–42.10.1136/annrheumdis-2016-209707Search in Google Scholar PubMed

[38] Tiwari S, Dwivedi H, Kymonil KM, Saraf SA. Urate crystal degradation for treatment of gout: a nanoparticulate combination therapy approach. Drug Delivery Transl Res. 2015;5:219–30.10.1007/s13346-015-0219-1Search in Google Scholar PubMed

[39] Parekh B, Vasant S, Tank K, Raut A, Vaidya A, Joshi M. In vitro growth and inhibition studies of monosodium urate monohydrate crystals by different herbal extracts. Am J infect Dis. 2009;5:225–30.10.3844/ajidsp.2009.225.230Search in Google Scholar

[40] Ansari SS, Diño PH, Castillo AL, Santiago LA. Antioxidant activity, xanthine oxidase inhibition and acute oral toxicity of Dillenia philippinensis Rolfe (Dilleniaceae) leaf extract. J Pharm Pharmacogn Res. 2021;9:846–58.10.56499/jppres21.1106_9.6.846Search in Google Scholar

[41] Ihsan J, Farooq M, Khan MA, Ghani M, Shah LA, Saeed S, et al. Synthesis, characterization, and biological screening of metal nanoparticles loaded gum acacia microgels. Microsc Res Tech. 2021;84:1673–84.10.1002/jemt.23726Search in Google Scholar PubMed

[42] Aziz H, Saeed A, Khan MA, Afridi S, Jabeen F, Hashim M. Novel N‐acyl‐1H‐imidazole‐1‐carbothioamides: Design, synthesis, biological and computational studies. Chem Biodivers. 2020;17:e1900509.10.1002/cbdv.201900509Search in Google Scholar PubMed

[43] Aziz H, Saeed A, Khan MA, Afridi S, Jabeen F. Synthesis, characterization, antimicrobial, antioxidant and computational evaluation of N-acyl-morpholine-4-carbothioamides. Mol Diversity. 2021;25:763–76.10.1007/s11030-020-10054-wSearch in Google Scholar PubMed

[44] Farooq M, Ihsan J, M K Mohamed R, Khan MA, Rehman TU, Ullah H, et al. Highly biocompatible formulations based on Arabic gum Nano composite hydrogels: Fabrication, characterization, and biological investigation. Int J Biol Macromol. 2022;209:59–69.10.1016/j.ijbiomac.2022.03.162Search in Google Scholar PubMed

[45] Roy K, Srivastwa AK, Ghosh CK. Anticoagulant, thrombolytic and antibacterial activities of Euphorbia acruensis latex-mediated bioengineered silver nanoparticles. Green Process Synth. 2019;8:590–9.10.1515/gps-2019-0029Search in Google Scholar

[46] Gunti L, Dass RS, Kalagatur NK. Phytofabrication of selenium nanoparticles from Emblica officinalis fruit extract and exploring its biopotential applications: antioxidant, antimicrobial, and biocompatibility. Front Microbiol. 2019;10:931.10.3389/fmicb.2019.00931Search in Google Scholar PubMed PubMed Central

[47] Xiong J, Wang Y, Xue Q, Wu X. Synthesis of highly stable dispersions of nanosized copper particles using L-ascorbic acid. Green Chem. 2011;13:900–4.10.1039/c0gc00772bSearch in Google Scholar

[48] Shin Y, Blackwood JM, Bae I-T, Arey BW, Exarhos GJ. Synthesis and stabilization of selenium nanoparticles on cellulose nanocrystal. Mater Lett. 2007;61:4297–300.10.1016/j.matlet.2007.01.091Search in Google Scholar

[49] Avendaño R, Chaves N, Fuentes P, Sánchez E, Jiménez JI, Chavarría M. Production of selenium nanoparticles in Pseudomonas putida KT2440. Sci Rep. 2016;6:37155.10.1038/srep37155Search in Google Scholar PubMed PubMed Central

[50] Sajadi SM, Nasrollahzadeh M, Maham M. Aqueous extract from seeds of Silybum marianum L. as a green material for preparation of the Cu/Fe3O4 nanoparticles: a magnetically recoverable and reusable catalyst for the reduction of nitroarenes. J Colloid Interface Sci. 2016;469:93–8.10.1016/j.jcis.2016.02.009Search in Google Scholar PubMed

[51] Sarkar J, Dey P, Saha S, Acharya K. Mycosynthesis of selenium nanoparticles. Micro Nano Lett. 2011;6:599–602.10.1049/mnl.2011.0227Search in Google Scholar

[52] Dehlin M, Jacobsson L, Roddy E. Global epidemiology of gout: prevalence, incidence, treatment patterns and risk factors. Nat Rev Rheumatol. 2020;16:380–90.10.1038/s41584-020-0441-1Search in Google Scholar PubMed

[53] Liang J, Jiang Y, Huang Y, Song W, Li X, Huang Y. The comparison of dyslipidemia and serum uric acid in patients with gout and asymptomatic hyperuricemia: a cross-sectional study. Lipids Health Dis. 2020;19:1–7.10.1186/s12944-020-1197-ySearch in Google Scholar PubMed PubMed Central

[54] Mehmood A, Ishaq M, Zhao L, Safdar B, Rehman AU, Munir M, et al. Natural compounds with xanthine oxidase inhibitory activity: A review. Chem Biol & drug Des. 2019;93:387–418.10.1111/cbdd.13437Search in Google Scholar PubMed

[55] Gutteridge JM. Biological origin of free radicals, and mechanisms of antioxidant protection. Chemico-Biol Interact. 1994;91:133–40.10.1016/0009-2797(94)90033-7Search in Google Scholar PubMed

[56] Bano I, Malhi M, Khatri P, Soomro SA, Sajjad H, Leghari A, et al. Effect of dietary selenium yeast supplementation on morphology and antioxidant status in testes of young goat. Pak J Zool. 2019;51:979.10.17582/journal.pjz/2019.51.3.979.988Search in Google Scholar

[57] Chen T, Wong Y-S. In vitro antioxidant and antiproliferative activities of selenium-containing phycocyanin from selenium-enriched Spirulina platensis. J Agric Food Chem. 2008;56:4352–8.10.1021/jf073399kSearch in Google Scholar PubMed

[58] Kainat MA, Khan MA, Ali F, Faisal S, Rizwan M, Hussain Z, et al. Exploring the therapeutic potential of Hibiscus rosa sinensis synthesized cobalt oxide (Co3O4-NPs) and magnesium oxide nanoparticles (MgO-NPs). Saudi J Biol Sci. 2021;28:5157–67.10.1016/j.sjbs.2021.05.035Search in Google Scholar PubMed PubMed Central

[59] Zhai X, Zhang C, Zhao G, Stoll S, Ren F, Leng X. Antioxidant capacities of the selenium nanoparticles stabilized by chitosan. J Nanobiotechnol. 2017;15:1–12.10.1186/s12951-016-0243-4Search in Google Scholar PubMed PubMed Central

[60] Kong H, Yang J, Zhang Y, Fang Y, Nishinari K, Phillips GO. Synthesis and antioxidant properties of gum arabic-stabilized selenium nanoparticles. Int J Biol macromolecules. 2014;65:155–62.10.1016/j.ijbiomac.2014.01.011Search in Google Scholar PubMed

[61] Faisal S, Khan MA, Jan H, Shah SA, Abdullah S, Shah S, et al. Edible mushroom (Flammulina velutipes) as biosource for silver nanoparticles: from synthesis to diverse biomedical and environmental applications. Nanotechnology. 2020;32:065101.10.1088/1361-6528/abc2ebSearch in Google Scholar PubMed

[62] Khurana A, Tekula S, Saifi MA, Venkatesh P, Godugu C. Therapeutic applications of selenium nanoparticles. Biomed Pharmacother. 2019;111:802–12.10.1016/j.biopha.2018.12.146Search in Google Scholar PubMed

[63] Pyrzynska K, Sentkowska A. Biosynthesis of selenium nanoparticles using plant extracts. J Nanostruct Chem. 2022;12:467–80.10.1007/s40097-021-00435-4Search in Google Scholar

[64] Boroumand S, Safari M, Shaabani E, Shirzad M, Faridi-Majidi R. Selenium nanoparticles: synthesis, characterization and study of their cytotoxicity, antioxidant and antibacterial activity. Mater Res Express. 2019;6:0850d8.10.1088/2053-1591/ab2558Search in Google Scholar

[65] Nabi F, Arain MA, Hassan F, Umar M, Rajput N, Alagawany M, et al. Nutraceutical role of selenium nanoparticles in poultry nutrition: a review. World’s Poult Sci J. 2020;76:459–71.10.1080/00439339.2020.1789535Search in Google Scholar

[66] Cruz LY, Wang D, Liu J. Biosynthesis of selenium nanoparticles, characterization and X-ray induced radiotherapy for the treatment of lung cancer with interstitial lung disease. J Photochem Photobiol B: Biol. 2019;191:123–7.10.1016/j.jphotobiol.2018.12.008Search in Google Scholar PubMed

[67] Riaz M, Sharafat U, Zahid N, Ismail M, Park J, Ahmad B, et al. Synthesis of Biogenic Silver Nanocatalyst and their Antibacterial and Organic Pollutants Reduction Ability. ACS Omega. 2022;7:14723–34.10.1021/acsomega.1c07365Search in Google Scholar PubMed PubMed Central

[68] Kokila K, Elavarasan N, Sujatha V. Diospyros montana leaf extract-mediated synthesis of selenium nanoparticles and their biological applications. N J Chem. 2017;41:7481–90.10.1039/C7NJ01124ESearch in Google Scholar

[69] Sharma G, Sharma AR, Bhavesh R, Park J, Ganbold B, Nam J-S, et al. Biomolecule-mediated synthesis of selenium nanoparticles using dried Vitis vinifera (raisin) extract. Molecules. 2014;19:2761–70.10.3390/molecules19032761Search in Google Scholar PubMed PubMed Central

[70] Ismail M, Gul S, Khan M, Khan MA, Asiri AM, Khan SB. Green synthesis of zerovalent copper nanoparticles for efficient reduction of toxic azo dyes Congo red and methyl orange. Green Process Synth. 2019;8:135–43.10.1515/gps-2018-0038Search in Google Scholar

[71] Ismail M, Gul S, Khan M, Khan MA, Asiri AM, Khan SB. Medicago polymorpha-mediated antibacterial silver nanoparticles in the reduction of methyl orange. Green Process Synth. 2019;8:118–27.10.1515/gps-2018-0030Search in Google Scholar

[72] Ismail M, Khan MI, Akhtar K, Khan MA, Asiri AM, Khan SB. Biosynthesis of silver nanoparticles: A colorimetric optical sensor for detection of hexavalent chromium and ammonia in aqueous solution. Phys E. 2018;103:367–76.10.1016/j.physe.2018.06.015Search in Google Scholar

[73] Gavin Jr R, Risemberg S, Rice SA. Spectroscopic properties of polyenes. I. The lowest energy allowed singlet‐singlet transition for cis‐and trans‐1, 3, 5‐hexatriene. J Chem Phys. 1973;58:3160–5.10.1063/1.1679637Search in Google Scholar

[74] Saraiva GD, Lima JA, de Sousa FF, Freire PTC, Filho JM, Filho AGS. Temperature dependent Raman scattering study of l-ascorbic acid. Vib Spectrosc. 2011;55:101–6.10.1016/j.vibspec.2010.09.006Search in Google Scholar

[75] Srivastava P, Braganca JM, Kowshik M. In vivo synthesis of selenium nanoparticles by Halococcus salifodinae BK18 and their anti‐proliferative properties against HeLa cell line. Biotechnol Prog. 2014;30:1480–7.10.1002/btpr.1992Search in Google Scholar PubMed

[76] Kalishwaralal K, Jeyabharathi S, Sundar K, Muthukumaran A. A novel one-pot green synthesis of selenium nanoparticles and evaluation of its toxicity in zebrafish embryos, Artificial cells. Nanomed, Biotechnol. 2016;44:471–7.10.3109/21691401.2014.962744Search in Google Scholar PubMed

[77] Jessica M, Oppedisano F, Gratteri S, Muscoli C, Mollace V. Regulation of uric acid metabolism and excretion. Int J Cardiol. 2016;213:8–14.10.1016/j.ijcard.2015.08.109Search in Google Scholar PubMed

[78] Zitt E, Fischer A, Lhotta K, Concin H, Nagel G. Sex-and age-specific variations, temporal trends and metabolic determinants of serum uric acid concentrations in a large population-based Austrian cohort. Sci Rep. 2020;10:7578.10.1038/s41598-020-64587-zSearch in Google Scholar PubMed PubMed Central

[79] Jansen TTA, Reinders M, Van Roon E, Brouwers J. Benzbromarone withdrawn from the European market: Another case of absence of evidence is evidence of absence?. Clin Exp Rheumatol. 2004;22:651.Search in Google Scholar

[80] Ramasamy SN, Korb-Wells CS, Kannangara DR, Smith MW, Wang N, Roberts DM, et al. Allopurinol hypersensitivity: a systematic review of all published cases, 1950–2012. Drug Saf. 2013;36:953–80.10.1007/s40264-013-0084-0Search in Google Scholar PubMed

[81] Raut AA, Sunder S, Sarkar S, Pandita NS, Vaidya AD. Preliminary study on crystal dissolution activity of Rotula aquatica, Commiphora wightii and Boerhaavia diffusa extracts. Fitoterapia. 2008;79:544–7.10.1016/j.fitote.2008.06.001Search in Google Scholar PubMed

[82] Kim J, Kim W, Hyun J, Lee J, Kwon J, Seo C, et al. Salvia plebeia extract inhibits xanthine oxidase activity in vitro and reduces serum uric acid in an animal model of hyperuricemia. Planta Med. 2017;83:1335–41.10.1055/s-0043-111012Search in Google Scholar PubMed

[83] Dong Y, Chi Y, Lin X, Zheng L, Chen L, Chen G. Nano-sized platinum as a mimic of uricase catalyzing the oxidative degradation of uric acid. Phys Chem Chem Phys. 2011;13:6319–24.10.1039/c0cp01759kSearch in Google Scholar PubMed

[84] Serrano JL, Lopes D, Reis MJ, Boto RE, Silvestre S, Almeida P. Bis-thiobarbiturates as promising xanthine oxidase inhibitors: synthesis and biological evaluation. Biomedicines. 2021;9:1443.10.3390/biomedicines9101443Search in Google Scholar PubMed PubMed Central

[85] Al-Hakeim HK, Kareem MM, Grulke EA. Synthesis a new magnetic nanoparticles and study the interaction with xanthine oxidase. Am J Nanomater. 2014;2:13–20.Search in Google Scholar

[86] Sindhi V, Gupta V, Sharma K, Bhatnagar S, Kumari R, Dhaka N. Potential applications of antioxidants–A review. J Pharm Res. 2013;7:828–35.10.1016/j.jopr.2013.10.001Search in Google Scholar

[87] Yasin MT, Ali Y, Ahmad K, Ghani A, Amanat K, Basheir MM, et al. Alkaline lipase production by novel meso-tolerant psychrophilic Exiguobacterium sp. strain (AMBL-20) isolated from glacier of northeastern Pakistan. Arch Microbiol. 2021;203:1309–20.10.1007/s00203-020-02133-1Search in Google Scholar PubMed

[88] Muhammad W, Khan MA, Nazir M, Siddiquah A, Mushtaq S, Hashmi SS, et al. Papaver somniferum L. mediated novel bioinspired lead oxide (PbO) and iron oxide (Fe2O3) nanoparticles: In-vitro biological applications, biocompatibility and their potential towards HepG2 cell line. Mater Sci Eng C. 2019;103:109740.10.1016/j.msec.2019.109740Search in Google Scholar PubMed

[89] Altaf F, Wu S, Kasim V. Role of fibrinolytic enzymes in anti-thrombosis therapy. Front Mol Biosci. 2021;8:680397.10.3389/fmolb.2021.680397Search in Google Scholar PubMed PubMed Central

[90] Lateef A, Folarin BI, Oladejo SM, Akinola PO, Beukes LS, Gueguim-Kana EB. Characterization, antimicrobial, antioxidant, and anticoagulant activities of silver nanoparticles synthesized from Petiveria alliacea L. leaf extract. Prep Biochem Biotechnol. 2018;48:646–52.10.1080/10826068.2018.1479864Search in Google Scholar PubMed

[91] Ajarem JS, Maodaa SN, Allam AA, Taher MM, Khalaf M. Benign synthesis of cobalt oxide nanoparticles containing red algae extract: antioxidant, antimicrobial, anticancer, and anticoagulant activity. J Clust Sci. 2022;33:717–28.10.1007/s10876-021-02004-9Search in Google Scholar

[92] Jin NZ, Gopinath SC. Potential blood clotting factors and anticoagulants. Biomed Pharmacother. 2016;84:356–65.10.1016/j.biopha.2016.09.057Search in Google Scholar PubMed

[93] Alhawiti AS. Citric acid-mediated green synthesis of selenium nanoparticles: antioxidant, antimicrobial, and anticoagulant potential applications. Biomass Convers Biorefin. 2022. 10.1007/s13399-022-02798-2 Search in Google Scholar PubMed PubMed Central

[94] Lien C-W, Chen Y-C, Chang H-T, Huang C-C. Logical regulation of the enzyme-like activity of gold nanoparticles by using heavy metal ions. Nanoscale. 2013;5:8227–34.10.1039/c3nr01836aSearch in Google Scholar PubMed

[95] Vemuri SK, Banala RR, Mukherjee S, Uppula P, Gpv S, AV GR, et al. Novel biosynthesized gold nanoparticles as anti-cancer agents against breast cancer: Synthesis, biological evaluation, molecular modelling studies. Mater Sci Eng: C. 2019;99:417–29.10.1016/j.msec.2019.01.123Search in Google Scholar PubMed

[96] Li T, Yuan D, Yuan J. Antithrombotic drugs – pharmacology and perspectives. Coron Artery Dis: Ther Drug Discovery. 2020;1177:101–31.10.1007/978-981-15-2517-9_4Search in Google Scholar PubMed

[97] Nirmala C, Sridevi M, Aishwarya A, Perara R, Sathiyanarayanan Y. Pharmacological prospects of morin conjugated selenium nanoparticles – evaluation of antimicrobial, antioxidant, thrombolytic, and anticancer activities. BioNanoScience. 2023;13:2469–82.10.1007/s12668-023-01116-ySearch in Google Scholar PubMed PubMed Central

[98] Chakraborty D, Chauhan P, Alex SA, Chaudhary S, Ethiraj KR, Chandrasekaran N, et al. Comprehensive study on biocorona formation on functionalized selenium nanoparticle and its biological implications. J Mol Liq. 2018;268:335–42.10.1016/j.molliq.2018.07.070Search in Google Scholar

[99] Hashem AH, Khalil AMA, Reyad AM, Salem SS. Biomedical applications of mycosynthesized selenium nanoparticles using Penicillium expansum ATTC 36200. Biol Trace Elem Res. 2021;199:1–11.10.1007/s12011-020-02506-zSearch in Google Scholar PubMed

[100] Alizadeh SR, Abbastabar M, Nosratabadi M, Ebrahimzadeh MA. High antimicrobial, cytotoxicity, and catalytic activities of biosynthesized selenium nanoparticles using Crocus caspius extract. Arab J Chem. 2023;16:104705.10.1016/j.arabjc.2023.104705Search in Google Scholar

[101] Puri A, Patil S. Tinospora cordifolia stem extract-mediated green synthesis of selenium nanoparticles and its biological applications. Pharmacog Res. 2022;14:289–96.10.5530/pres.14.3.42Search in Google Scholar

[102] Alagesan V, Venugopal S. Green synthesis of selenium nanoparticle using leaves extract of withania somnifera and its biological applications and photocatalytic activities. Bionanoscience. 2019;9:105–16.10.1007/s12668-018-0566-8Search in Google Scholar

[103] Iqbal MS, Abbas K, Qadir MI. Synthesis, characterization and evaluation of biological properties of selenium nanoparticles from Solanum lycopersicum. Arab J Chem. 2022;15:103901.10.1016/j.arabjc.2022.103901Search in Google Scholar

Received: 2023-08-22
Accepted: 2023-12-04
Published Online: 2024-03-05

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

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

Downloaded on 15.5.2024 from https://www.degruyter.com/document/doi/10.1515/gps-2023-0158/html
Scroll to top button