[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Article in Journal
Challenges to Accident Prevention for High-Speed Vessels Used in the Brazilian Amazon
Previous Article in Journal
A Potential Technology for Improving the Quality of Lake Water Environments in Cold Zones: A Case Study from Lake Hulun, China
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sustainable Green Synthesis of ZnO Nanoparticles from Bromelia pinguin L.: Photocatalytic Properties and Their Contribution to Urban Habitability

by
Manuel de Jesus Chinchillas-Chinchillas
1,
Horacio Edgardo Garrafa Galvez
2,
Victor Manuel Orozco Carmona
3,
Hugo Galindo Flores
1,
Jose Belisario Leyva Morales
4,
Mizael Luque Morales
5,
Mariel Organista Camacho
5,* and
Priscy Alfredo Luque Morales
5,*
1
Departamento de Ingeniería y Tecnología, Universidad Autónoma de Occidente (UAdeO), Los Mochis C.P. 81048, Sinaloa, Mexico
2
Facultad de Ingeniería Mochis, Universidad Autónoma de Sinaloa (UAS), Los Mochis C.P. 81223, Sinaloa, Mexico
3
Departamento de Metalurgia e Integridad Estructural, Centro de Investigación en Materiales Avanzados (CIMAV), Chihuahua C.P. 31136, Chihuahua, Mexico
4
Instituto de Ciencias Básicas e Ingeniería, Universidad Autónoma del Estado de Hidalgo (UAEH), Mineral de la Reforma C.P. 42184, Hidalgo, Mexico
5
Facultad de Ingeniería, Arquitectura y Diseño, Universidad Autónoma de Baja California (UABC), Ensenada C.P. 22860, Baja California, Mexico
*
Authors to whom correspondence should be addressed.
Sustainability 2024, 16(23), 10745; https://doi.org/10.3390/su162310745
Submission received: 9 November 2024 / Revised: 2 December 2024 / Accepted: 5 December 2024 / Published: 7 December 2024
Figure 1
<p>ATR-IR analysis of the nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L.</p> ">
Figure 2
<p>Morphological analysis of the nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L. (<b>a</b>,<b>b</b>) Micrographs of BP 1%-ZnO, (<b>c</b>) size distribution of BP 1%-ZnO, (<b>d</b>,<b>e</b>) Micrographs of BP 2%-ZnO, (<b>f</b>) size distribution of BP 2%-ZnO, (<b>g</b>,<b>h</b>) Micrographs of BP 4%-ZnO and (<b>i</b>) size distribution of BP 4%-ZnO.</p> ">
Figure 3
<p>XRD spectra of the nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L. (<b>a</b>) BP 1%-ZnO nanoparticles, (<b>b</b>) BP 2%-ZnO nanoparticles and (<b>c</b>) BP 4%-ZnO nanoparticles.</p> ">
Figure 4
<p>TGA/DSC results of the nanoparticles synthesized using 1%, 2%, and 4% of <span class="html-italic">Bromelia pinguin</span> L.</p> ">
Figure 5
<p>BET analysis of the nanoparticles synthesized using 1%, 2%, and 4% of <span class="html-italic">Bromelia pinguin</span> L.</p> ">
Figure 6
<p>UV–Vis spectra of nanoparticles synthesized using 1%, 2%, and 4% of <span class="html-italic">Bromelia pinguin</span> L.</p> ">
Figure 7
<p>Band gaps of the nanoparticles synthesized using 1%, 2%, and 4% of <span class="html-italic">Bromelia pinguin</span> L. calculated with the help of the TAUC model.</p> ">
Figure 8
<p>Formation mechanism of nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L.</p> ">
Figure 9
<p>Photocatalytic activity of nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L. (<b>a</b>) Degradation of MB under solar radiation; (<b>b</b>) degradation of MB under UV radiation; (<b>c</b>) MO degradation under solar radiation; (<b>d</b>) degradation of MO under UV radiation; (<b>e</b>) degradation of RhB under solar radiation; and (<b>f</b>) degradation of RhB under UV radiation.</p> ">
Figure 10
<p>Proposed photocatalytic degradation mechanism of ZnO nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L.</p> ">
Versions Notes

Abstract

:
Aguama (Bromelia pinguin L.), a plant belonging to the Bromeliaceae family, possesses a rich content of organic compounds historically employed in traditional medicine. This research focuses on the sustainable synthesis of ZnO nanoparticles via an eco-friendly route using 1, 2, and 4% of Aguama peel extract. This method contributes to environmental sustainability by reducing the use of hazardous chemicals in nanoparticle production. The optical properties, including the band gap, were determined using the TAUC model through Ultraviolet–Visible Spectroscopy (UV–Vis). The photocatalytic activity was evaluated using three widely studied organic dyes (methylene blue, methyl orange, and rhodamine B) under both solar and UV radiation. The results demonstrated that the ZnO nanoparticles, characterized by a wurtzite-type crystalline structure and particle sizes ranging from 68 to 76 nm, exhibited high thermal stability and band gap values between 2.60 and 2.91 eV. These nanoparticles successfully degraded the dyes completely, with methylene blue degrading in 40 min, methyl orange in 70 min, and rhodamine B in 90 min. This study underscores the potential of Bromelia pinguin L. extract in advancing sustainable nanoparticle synthesis and its application in environmental remediation through efficient photocatalysis.

1. Introduction

Worldwide, water is used for all economic activities carried out by humans, including agriculture, livestock, the mining sector, various manufacturing sectors, construction, and electricity. As a result of these, pollutants are generated, such as pesticides in farming and dyes produced by the textile sector which are dispersed in large quantities to nearby ecosystems, both terrestrial and aquatic, negatively affecting the biota that inhabits them, as well as indirectly affecting human health [1,2,3]. Organic dyes have emerged as highly significant pollutants in recent years due to the discharge of effluents from various industries, including food, cosmetics, pharmaceuticals, plastics, paints, and textiles, among others. These dyes pose a substantial environmental challenge as they are non-biodegradable and tend to bioaccumulate in organisms, leading to severe health issues such as cancer, allergies, dermatitis, genetic mutations, and skin irritation. Furthermore, they disrupt aquatic ecosystems by hindering photosynthesis through sunlight blockage [4,5,6,7]. Due to the reasons, a wide range of treatments have been used for the total or partial elimination of this type of compound, highlighting coagulation, membrane filtration, ozonation, flocculation, oxidation (photochemistry and electrochemistry), sonochemical decomposition, sedimentation, photocatalysis (homogeneous and heterogeneous), among others. Of these, photocatalytic processes have shown to have high efficiency in the degradation of recalcitrant contaminants. Their foundation is based on the induction of highly oxidative and reductive processes through the generation of hydroxyl radicals (OH) and superoxides (O2−) that result from photocatalytic activity when a semiconductor interacts with a source of photons (visible light or UV) in an aqueous medium, causing contaminants to mineralize and generate non-toxic byproducts [8]. There are many semiconductor materials used for this purpose, among which the metal oxides of copper, iron, titanium, manganese, nickel, antimony, and zinc oxide stand out [9,10,11,12,13]. The latter, in its nanostructured form, is a semiconductor with great potential for being a low-cost material, which has high photocatalytic efficiency related to its forbidden energy gap or band gap, crystallinity, particle size, and contact area, making it a photocatalytic agent with potential application in the bioremediation of water contaminated with organic industrial waste [14,15,16]. Nanoparticles can be produced through different methods, generally categorized as “top-down” and “bottom-up”. Physical techniques encompass mechanical milling, laser ablation, sputter deposition, vacuum vapor deposition, lithography, pulsed wire discharge (PWD), and arc discharge methods; while within the chemicals, we find microemulsion, photochemical reduction, pyrolysis, electrochemical methods, thermal decomposition, sol-gel, and microwave-assisted combustion method [17,18]. However, several of these methods employ the use of hazardous chemicals, complex processes with significant costs, elevated operating pressures and temperatures with high energy consumption, and/or the generation of secondary toxic waste. Biological synthesis of nanoparticles can also be achieved with various microorganisms, including bacteria, algae, actinomycetes, viruses, and yeasts that present some very marked disadvantages such as slow growth rates, multiple production steps requiring significant time, and a restricted range of achievable sizes and shapes. On the other hand, macroscopic organisms such as plants (leaves, stem, root, fruit, etc.) are also used through the so-called green synthesis, which is based on the use of extracts (used as reducing and stabilizing), as it replaces the hazardous chemicals used in conventional methods, in addition to being efficient, fast, and economical [8,17,18,19,20]. Bromelia pinguin L., a plant widely distributed in the arid regions of Mexico, Venezuela, Costa Rica, and the Caribbean, represents an accessible and underutilized natural resource [21]. This resource is particularly appealing due to its high availability in areas where it is generated as agricultural waste, allowing its use without competing with other applications [22,23]. Moreover, its utilization aids in the valorization of organic waste, aligning with the core of sustainability and green chemistry. The peel of this plant is rich in functional compounds, including flavonoids, phenolic compounds, tannins, and proteins [24,25]. These compounds serve as excellent reducing and stabilizing agents in nanoparticle synthesis, enabling an eco-friendly route for producing materials with enhanced properties. In particular, their unique chemical composition allows the formation of nanoparticles with reduced sizes and greater stability, thereby improving their applicability in photocatalytic processes. The remediation of contaminated water not only addresses environmental challenges but also plays a vital role in enhancing urban habitability. Clean water resources are essential for maintaining public health, supporting sustainable development, and improving the quality of life in densely populated urban areas. This research highlights how the green synthesis of ZnO nanoparticles utilizing Bromelia pinguin L. aligns with sustainability objectives through the efficient photocatalytic degradation of organic pollutants. ZnO nanoparticles were synthesized using green methods with extracts derived from the shell of aguama (Bromelia pinguin L.), and the impact of different extract concentrations (1%, 2%, and 4%) on the properties of the nanoparticles was analyzed. Furthermore, the synthesized ZnO nanoparticles were employed to degrade three industrial dyes—methylene blue (MB), methyl orange (MO), and rhodamine B (RhB)—under both solar and UV radiation.

2. Materials and Methods

Materials used in the synthesis. In the biosynthesis, the following were used: Aguama shell (Bromelia pinguin L.), donated by the Health Sciences Laboratory of the Autonomous University of the West, Culiacán Campus; Zinc nitrate [Zn(NO3)2·6H2O] at 98% purity, acquired from Sigma-Aldrich (produced by MilliporeSigma, a subsidiary of Merck KGaA, headquartered in Darmstadt, Germany) as a precursor salt; and deionized water, commercially obtained from Sumilab S.A. de C.V., headquartered in Mazatlán, Sinaloa, Mexico. The dyes used were MB (373.9 g/mol), MO (327.34 g/mol), and RhB (479.01 g/mol), all obtained from Sigma-Aldrich.
Extract obtention and Biosynthesis of ZnO nanoparticles. The aqueous medium for synthesizing nanoparticles through green synthesis was an extract obtained from the shell of aguama (Bromelia pinguin L.) with different percentages. To prepare the extract, the aguama shells were cut and dried at ambient temperature for 48 h. The dried shells were subsequently milled into a fine powder. Three different samples were weighed, one of 0.5 g (1%), 1 g (2%), and 2 g (4%), which were identified as BP 1%-ZnO, BP 2%-ZnO, and BP 4%-ZnO, respectively, where BP indicates the use of Bromelia pinguin L. extract in the synthesis process. This powdered material was added to beakers containing 50 mL of deionized water and stirred for 2 h at ambient temperature. After that time, it was filtered using a vacuum pump to eliminate the remains of shells in the solution, and in this way different percentages of natural extracts of Bromelia pinguin L. were obtained. The green synthesis process began by adding 2 g of ZnO precursor to the three different percentages of extracts obtained. The three solutions were immediately stirred for one hour at room temperature, and then placed in a thermal bath (bain-marie) at 60 °C for 11 h. Finally, the obtained material was transferred to porcelain capsules and heated at 400 °C for one hour. In this way, ZnO nanoparticles were obtained with extracts of Bromelia pinguin L.
Characterization of nanoparticles. To determine the properties of the biosynthesized nanoparticles, several characterization techniques were employed. ATR-IR analysis was conducted using. The PerkinElmer equipment (brand equipment) operating in the range of 4500 to 350 cm−1 was manufactured by PerkinElmer, Inc., headquartered in Waltham, Massachusetts, United States. XRD measurements were carried out with a Bruker-D2 Phase instrument, operating at 30 kV and 10 mA, equipped with a Cu-Kα radiation source (λ = 1.5406 Å). The instrument was manufactured by Bruker Corporation, headquartered in Billerica, Massachusetts, United States. For UV–Vis spectroscopy analysis, a Perkin Elmer Lambda 365 model with a scanning speed of 600 nm/min was manufactured by PerkinElmer, Inc., headquartered in Waltham, Massachusetts, Unit-ed States. These three analyses were conducted at the UABC, Ensenada Campus. Morphological analysis was conducted using a JEOL JSM-6310LV high-resolution SEM at 8 mm and 5 kV. This equipment was manufactured by JEOL Ltd., headquartered in Akishima, Tokyo, Japan. Thermal analysis was performed using a TA Instruments SDT Q600 unit (up to 800 °C with 10 °C/min) for TGA/DSC. This equipment was manufactured by TA Instruments, headquartered in New Castle, Delaware, United States. For the BET method, a TriStar II 3020 instrument (N2 adsorption at 77 K) was employed. This equipment was manufactured by Micromeritics Instrument Corporation, headquartered in Norcross, Georgia, United States. These characterization studies were conducted at the Advanced Materials Research Center (CIMAV-Chihuahua).
Photocatalytic test. To assess the photocatalytic degradation of nanoparticles synthesized using Bromelia pinguin L. extract, it was first necessary to prepare different contaminants commonly found in water. Three industrial dyes (MB, MO, and RhB) were dissolved in 50 mL of deionized water, each at a concentration of 15 ppm. Photolysis of the dyes was carried out under solar radiation and UV radiation, with measurements taken every 30 min for 4 h. Furthermore, 50 mg of nanoparticles synthesized with varying concentrations of Bromelia pinguin L. extracts were added to the dye solutions. They were subsequently shaken (in darkness) for 30 min before their photocatalytic performance was analyzed under solar and UV radiation using reactors (The Polaris UV-1C system, equipped with 10 W lamps, was manufactured by Polaris Scientific UV, headquartered in Chino Hills, California, United States). Measurements of the concentration of industrial dyes were evaluated every 10 min using the UV–Vis spectrophotometer equipment until complete elimination of the three dyes was observed.

3. Results

ATR-IR. The ATR-IR analysis of the nanoparticles biosynthesized using Bromelia pinguin L. is observed in Figure 1. Three areas are shown, represented by colors. The first area, represented in blue, corresponds to the vibrations of the O-H groups, indicating the adsorption of water molecules in all three samples analyzed [26]. The green area represents the vibrational modes of the molecules originating from the natural extract employed during the biosynthesis process. Previous studies have indicated that certain organic molecules from the extract remain attached to the nanoparticle surface after synthesis, contributing to improved stabilization [27]. The three samples exhibit a vibrational band at 1380 cm−1, which represents the bending vibrations of C-H groups found in the phytochemicals of Bromelia pinguin L. extract, such as flavonoids and phenolic compounds [28]. As the proportion of aguama shell used in the biosynthesis increases, the vibrational band becomes broader. This behavior is attributed to the higher concentration of organic molecules functionalized onto the ZnO nanoparticle surface when greater amounts of extract are used. After calcination at 400 °C, a significant reduction in the intensity of these bands was observed, particularly in C-H vibrations. However, some signals, such as those at 1650 cm−1 (C=O) and 3300 cm−1 (O-H), persist. This suggests that certain organic compounds remain anchored to the ZnO nanoparticle surface due to strong interactions between functional groups and Zn atoms, as previously reported [26].
Finally, at 370 cm−1, the vibration band corresponding to Zn-O is observed [29,30]. Residual organic residues adhered to the nanoparticle surface can actively contribute to photocatalysis by acting as photosensitizers and enhancing energy transfer to ZnO. This mechanism promotes the generation of reactive species, such as OH⋅ and O 2 , thereby improving the photocatalytic efficiency of the material [31]. This analysis confirms that the biosynthesized nanoparticles contain organic molecules and are semiconductor in nature.
SEM-HR. The SEM technique is an effective method to obtain high-resolution images of nanoparticles. With this technique, information about nanomaterials such as their topography, surface, and shape of the nanoparticle, among others, can be known [32]. As seen in Figure 2, all the samples analyzed in this study present a similar morphology. The nanoparticles biosynthesized with different percentages of Bromelia pinguin L. extract showed a hemispherical shape. Other studies have presented a similar morphology of ZnO nanoparticles using natural extracts [33,34]. Furthermore, there is a relationship between the amount of extract used in biosynthesis and the particle size. When nanoparticles are synthesized with a small amount of Bromelia pinguin L. extract (ZnO_BP1%), displayed in Figure 2a,b, the nanoparticles have a size of approximately 76 nm (Figure 2c). By increasing the amount of extract (ZnO_BP2%), the particle size decreases, observing an average size of 72 nm (Figure 2d–f). Furthermore, the nanoparticles biosynthesized with 4% (ZnO_BP4%) decreased in size, presenting average sizes of 68 nm (Figure 2g,h). This reduction in size can be ascribed to the interactions between the organic components in the extract and the nanoparticle synthesis process. When they are in the process of nucleation and growth, these organic molecules hinder or saturate the system, preventing agglomeration, causing them to decrease in size [35]. The micrographs also show that the greater the amount of extract used, the greater the agglomeration. Previous studies have reported that using greater amounts of natural extract in biosynthesis causes a decrease in the particle size [36]. Additionally, through the EDS analysis, it can be noted that the three samples present signals of zinc (Zn), oxygen (O), and potassium (K). As reported in the literature, the fruit of Bromelia pinguin L. presents large amounts of potassium in its chemical structure, and therefore it was detected in the EDS analysis [25]. An increase in the extract concentration during biosynthesis corresponds to a rise in the potassium signal. A signal of 3.5%, 4.7%, and 8.2% is observed for the ZnO_BP1% (Figure 2c), ZnO_BP2% (Figure 2f), and ZnO_BP4% (Figure 2i) samples, respectively. This suggests that the organic molecules from the extract remain actively involved throughout the nanoparticle formation process.
XRD. XRD analysis was carried out on the ZnO nanoparticles biosynthesized with Bromelia pinguin L., to know the diffraction indices and be able to relate them to their crystallinity. As seen in Figure 3, various peaks were presented at 31.74, 34.40, 36.22, 47.49, 56.52, 62.81, and 67.89°. These peaks belong to the indices (100), (002), (101), (102), (110), (103), and (112), respectively. According to the JCPDS card No: 36-145, these diffraction patterns confirm that the nanoparticles exhibit a hexagonal wurtzite structure [37,38]. In contrast, the crystallite size of the nanoparticles of the three most intense peaks presented in the diffractogram [(100), (002), (101)] was obtained following the Debye–Scherrer procedure, using the equation D = 0.9 λ β cosθ [39,40]. When the crystallite size of the three most intense peaks was obtained, an average size was obtained. The results showed a crystallite size of 37.82, 14.95, and 11.8 nm corresponding to the BP 1%-ZnO, BP 2%-ZnO, and BP 4%-ZnO samples, respectively. The crystallite size coincides with the SEM analysis, where it was noted that by increasing the amount of extract used in the biosynthesis, the average diameter of the nanoparticles decreased. This is attributed to the organic molecules in the extract prevent the growth of the nanoparticles by acting as a barrier, preventing agglomeration.
TGA/DSC. The thermal behavior of the nanoparticles is shown in Figure 4. For a more detailed analysis, the thermograms were divided into three sections influenced by exothermic processes [41]. These sections comprise Section 1, which contains weight losses that belong to a combination of adsorbed water, chemically bound water, and low-weight free molecules that belong to organic components of Bromelia pinguin L. extracts [42]. In Section 2, the weight losses that comprise the crystallization of the ZnO nanoparticles [43] are presented in conjunction with the loss of the organic components of the Bromelia pinguin L. extracts that chemically interact with ZnO nanoparticles (this phenomenon has already been reported previously [44]). In the last section, Section 3, a third weight loss is observed that includes the calcination of organic species from Bromelia pinguin L. extracts of higher weight and free Zn+2 species that failed to form nanoparticles [45].
Through an exhaustive analysis of the results obtained, which are shown in Table 1, it can be observed that the BP 4%-ZnO nanoparticles have a greater total weight loss compared to BP 1%-ZnO and BP 2%-ZnO. This outcome aligns with expectations, as a greater concentration of Bromelia pinguin L. extract leads to an increased presence of organic species [46]. It is also possible to observe that by increasing the extract concentration, the stability of the species in Section 1 and Section 2 increases. This may be ascribed to the increased concentration of Bromelia pinguin L. extract used, which results in a greater abundance of organic species. The interaction between these species and the Zn2+ species gives way to a greater formation of ZnO nanoparticles.
BET. Figure 5 shows the results of BET analysis, indicating that the three types of synthesized nanoparticles present type IV isotherms in accordance with the IUPAC classification. This suggests that the three study samples (BP 1%-ZnO, BP 2%-ZnO, and BP 4%-ZnO) are mesoporous materials, consistent with results shown in previous research [47]. It is noticeable that for the three types of nanoparticles, as the relative pressure increases, the volume of adsorbed gas also increases. This is due to the fact that nanoparticles have sizes on the nanometer scale, and the surface area is much greater compared to micro materials and bulk materials [48]. The volume of adsorbed gas increases following this order: BP 1%-ZnO < BP 2%-ZnO < BP 4%-ZnO. An increased extract concentration results in to an increased presence of organic molecules acting as stabilizing agents, which inhibit agglomeration and growth, resulting in smaller nanoparticles and consequently larger surface areas [49]. The numerical data, summarized in Table 2, confirm these observations. By increasing the concentration of Bromelia pinguin L. extract used, the surface area increases, with BP 4%-ZnO being the nanoparticles with the highest value (21.98 m2/g). As mentioned above, the study samples are mesoporous, and when the pore volume was determined, something similar was observed. The nanoparticles biosynthesized with 4% have a larger pore volume than BP 1%-ZnO and BP 2%-ZnO. Additionally, Table 2 shows that pore size decreases with increasing extract concentration, a phenomenon attributed to the reduction in nanoparticle size [50].
UV–Vis and Band gap. The optical behavior through UV–Vis analysis of the ZnO nanoparticles obtained in this research is observed in Figure 6. The results revealed that the BP 1%-ZnO, BP 2%-ZnO, and BP 4%-ZnO samples exhibited characteristic spectra of ZnO nanoparticles, with absorption peaks observed at 374.5 nm, 371.2 nm, and 367.9 nm, respectively, for the samples synthesized with 1%, 2%, and 4% extract. These peaks are associated with electronic transitions between the valence band and the conduction band. Since no other band is present, it can be ensured that we have pure nanoparticles [51]. This peak can also be used to determine the band gap.
Using the UV–Vis results, the energy gap was determined through the TAUC model. This model analyzes the optical absorbance in relation to energy, following the principles of the Lambert–Beer law [52]. The TAUC model establishes a relationship between the band gap and photon energy [53]. Figure 7 illustrates that the ZnO nanoparticles obtained using 1%, 2%, and 4% of Bromelia pinguin L. extract exhibited lower band gap values compared to commercial ZnO (3.37 eV [54]), which were 2.91 eV, 2.89 eV, and 2.60 eV for BP 1%-ZnO, BP 2%-ZnO, and BP 4%-ZnO, respectively. Comparable results have been documented in the literature, such as the research by C.A. Soto-Robles et al., in 2019, where they presented ZnO band gap values of 2.86 eV, biosynthesizing the nanoparticles with extracts of Hibiscus sabdariffa [27]. Furthermore, in 2023, Sherwan M. et al. obtained values of 2.6 eV in ZnO nanoparticles biosynthesized with Pinus brutia leaf extract [55]. This effect arises from the direct interaction between the synthesized nanoparticles and the organic constituents of the Bromelia pinguin L. extracts. It is possible to obtain some defects in the nanoparticles such as oxygen vacancies [31]. Furthermore, organic molecules like flavonoids and polyphenols in the extract act as photosensitizers [56], causing the band gap to reduce its value. It is worth mentioning that the lower the band gap value, the less energy the semiconductor nanoparticles need to produce the energy jump of their electrons, which is beneficial for the application of photocatalysis.
Mechanism of formation of nanoparticles. Figure 8 shows the process of nanoparticle formation using Bromelia pinguin L. as a reducing and/or stabilizing agent. As described in the methodology, first, it is necessary to obtain the natural extract free of impurities (organic material). A specific quantity of zinc nitrate is added to the extract and mixed using magnetic stirring. The biosynthesis involves three key stages: hydrolysis, complexation, and decomposition. Organic molecules in the extract serve as reducing agents, donating electrons from carbonyl groups, leading to the formation of a Zn complex and the reduction of zero-valent Zn2+ ions. Finally, the polycondensation of the Zn complex into ZnO nanoparticles follows [57,58]. As the base of the extract is an aqueous solution, at this moment, a chemical hydrolysis reaction begins, where the original molecule of the metal precursor begins to decompose into smaller fragments [59]. Notably, the organic compounds in the extract can neutralize or bioreduce reactive oxygen species, free radicals, and metal chelates, facilitating the green synthesis process [60]. Subsequently, the solution is given a heating treatment at 60 °C, which causes an acceleration in the decomposition of zinc nitrate, causing a reduction to zinc hydroxide (as shown in Figure 8), and subsequently, nucleation begins, and the growth of the nanomaterial [20]. The next step in the synthesis of the nanoparticles is the heat treatment at 400 °C, where the material begins to crystallize. It is worth mentioning that the molecules of the extract remain interacting at all times, and it has been reported that these molecules become anchored on the surface of the material with interaction with the hydroxyl groups. These organic molecules prevent excessive growth of the material and manage to stabilize them [61,62]. As demonstrated by ATR-IR and EDS analyses, a significant portion of organic molecules remains even after the calcination process, contributing to a reduction in nanoparticle size (see Figure 2), an increase in thermal properties (see Figure 4), an increase in surface area (see Figure 5), and a decrease in the band gap (see Figure 7).
Photocatalytic degradation. The analysis of photocatalytic degradation was carried out under two different conditions: under solar radiation at a temperature of 34 °C and under UV radiation using 10 W reactors. Under these two conditions, the effect of nanoparticles biosynthesized with Bromelia pinguin L. on the degradation of organic pollutants was evaluated. The results are seen in Figure 9. Organic dyes used in various industries have aromatic groups in their structure, making them difficult to degrade naturally. The chromophores responsible for the color of the analyzed dyes include a thiazine ring in methylene blue (MB), an azo group (-N=N-) in methyl orange (MO), and a xanthene ring in rhodamine B (RhB). These chromophores are targeted by active species such as hydroxyl radicals (OH⋅) and superoxide anions ( O 2 . ), which break their π-conjugated bonds, leading to molecular degradation and decolorization. Anju Chanu in 2019 reported the elimination of methylene blue under UV radiation, achieving a degradation of less than 2% after two hours of exposure [63]. Similar results were found by Zaid Hamid in 2019, reporting the photolysis of 1% MB for 90 min [64]. On the other hand, the photolysis of MO under UV radiation was reported by Chao Han in 2023, presenting a degradation of approximately 5% in 90 min [65]. Furthermore, Alshamsi. et.al., in 2022, degraded RhB dye under the sun, and no degradation was observed after 300 min [66]. On the other hand, when a photocatalyst material is added, the degradation of pollutants is significantly accelerated. C.A. Soto-Robles et.al. in 2021 used ZnO nanoparticles biosynthesized with Justicia spicigera and degraded 90% of MB under UV radiation in 120 min [67]. Nguyen-Hong et al., in 2021, used Ce-ZnO nanoparticles through an extract of Hedyotis capitellata, managing to degrade 92% of the MO in 240 min under solar radiation [68], and Munir Ahmad et.al., in 2021, used Au-ZnO nanoparticles obtained through green synthesis with Carya illinoinensis extract, degrading 95% of RhB under UV radiation in 180 min [69]. In this work, good results were obtained compared to some reported in the literature. Using different percentages of natural extracts in the formation of nanoparticles allowed the photocatalytic activity to increase. As shown in Figure 9, the degradation of the contaminant increases when using more extract in all the study samples. The degradation in the dark of all the samples analyzed does not represent a significant change in the concentration of the three dyes. The elevated specific surface area of ZnO nanoparticles likely promotes the initial adsorption of dyes, improving their interaction with the photocatalyst during irradiation. However, experiments conducted in the dark showed no significant dye degradation, confirming that adsorption alone is insufficient and that photocatalysis is the primary mechanism. Figure 9a illustrates the photocatalytic activity of MB and it is observed that the three study samples were completely degraded. ZnO_BP1% degraded the MB in 120 min, ZnO_BP2% degraded the MB in 90 min, and ZnO_BP4% was the one that required the shortest time for degradation (50 min). In contrast, Figure 9b presents similar results under UV radiation; as the percentage of extract used increased, the photocatalytic activity increased. The difference is that when UV light was used, the degradation was accelerated due to the difference in energy between the two light sources. Sample ZnO_BP1% and ZnO_BP2% degraded MB in 80 min, and sample ZnO_BP4% achieved degradation in 40 min. In contrast, Figure 9c displays the findings of MO photocatalysis, where it was not possible to eliminate 100% of the contaminant with the three study samples. An effect of the amount of extract used in the biosynthesis is observed, since the degradations were 78%, 93%, and 98% for the samples ZnO_BP1%, ZnO_BP2%, and ZnO_BP4%, respectively, in 150 min of the study. When the analysis was carried out under UV radiation, this contaminant was completely degraded in the three samples (Figure 9d). ZnO_BP1% needed 120 min to achieve complete degradation, ZnO_BP2% degraded in 80 min, and ZnO_BP4% degraded in 70 min. Finally, the degradation of the RhB dye also had satisfactory results. Under sunlight, the three samples allowed the dye to be completely degraded, highlighting that the sample biosynthesized with 4% achieved degradation in 110 min (Figure 9e). Moreover, under UV radiation, this same sample degraded after 90 min (Figure 9f). As observed in the photocatalytic study, the ZnO nanoparticles biosynthesized with Bromelia pinguin L. yielded good results when degrading the dyes under solar radiation. This efficiency is attributed to the organic molecules in the natural extract, which facilitated a decrease in the nanoparticles’ band gap (2.6 eV for the BP 4%-ZnO sample), consequently enhancing their photocatalytic activity. It has been documented that smaller nanoparticle sizes result in a lower band gap, reducing the energy required for the electron’s energy jump and the generation of free radicals essential for the photocatalytic process [70]. Furthermore, nanoparticle size affects the surface area, and an increase in surface area results in more active sites, facilitating the photocatalytic degradation process [71]. These findings demonstrate that employing Bromelia pinguin L. in the biosynthesis of nanoparticles offers significant advantages for the efficient removal of organic dyes, a crucial aspect for their industrial applications. Moreover, the implications of such green synthesis methods extend beyond environmental benefits to positively impact urban habitability. By supporting cleaner water resources and reducing pollutants, these advances in photocatalytic degradation play a crucial role in enhancing quality of life in urban settings, aligning with the goals of sustainable development and improved habitability in densely populated areas. The efficient degradation of contaminants using ZnO nanoparticles synthesized with Bromelia pinguin L. significantly contributes to urban habitability. This green synthesis process avoids toxic chemicals, utilizing natural reducing and stabilizing agents like phenolic compounds and flavonoids, which enhance nanoparticle properties such as reduced size, lower band gap, and increased surface area. The nanoparticles effectively degraded methylene blue, methyl orange, and rhodamine B under solar and UV radiation, showcasing their potential for large-scale water treatment. By improving water quality and reducing environmental impact, these advancements promote sustainability, healthier ecosystems, and better quality of life in urban areas. This strategy highlights the economic and environmental benefits of renewable resources, fostering resilient, sustainable cities while supporting global water management goals.
Photocatalytic degradation mechanism. Figure 10 shows a proposal for the photocatalysis mechanism of ZnO nanoparticles based on studies by Luis Escobar in 2021 and K.V. Karthik in 2022, among others. The degradation mechanism begins when the ZnO nanoparticle (photocatalyst) biosynthesized with Bromelia pinguin L. is irradiated with energy in the form of light (solar or UV) with a wavelength equal to or exceeding the band gap energy of the semiconductor. When this happens, electrons are promoted from the valence band to the conduction band, generating electron (e) and hole (h+) pairs, with e residing in the conduction band and h+ in the valence band [72]. Therefore, the smaller the band gap of the ZnO nanoparticles, the lower the energy required to initiate this excitation and generate the electron–hole pairs. If these pairs avoid recombination and migrate to the nanoparticle surface, they participate in redox reactions with compounds present on the surface. H+ reacts by oxidizing water (H2O) and producing hydroxyl radicals (•OH) [31]. These radicals interact with MB, MO, and RhB molecules, degrading them and generating waste such as H2O and CO2. On the other hand, e reduces the oxygen (O2) adsorbed on the nanoparticle surface, forming superoxide radicals (O2−•). The O2−• radicals interact with the organic dye molecules, degrading them, generating H2O and CO2 as waste [14].

4. Conclusions

ZnO nanoparticles were successfully synthesized through green synthesis using Bromelia pinguin L. extract. The extract of this fruit allowed the modification of the properties of the nanomaterial. It was found that using a higher extract concentration in the biosynthesis caused a decrease in the size of the nanoparticles, a decrease in the crystallite dimensions, an increase in thermal resistance, increased surface area, and a reduced band gap. These modifications in properties significantly accelerated the photocatalytic degradation of three organic dyes. Higher extract percentages corresponded to improved photocatalytic performance. Under solar radiation, the ZnO_BP4% sample achieved 100% degradation of MB in 50 min, 98% degradation of MO in 150 min, and 100% degradation of RhB in 110 min. Similarly, under UV radiation, the ZnO_BP4% sample completely degraded all three dyes, with MB degrading in 40 min, MO in 70 min, and RhB in 90 min. The results obtained in this research are outstanding; they make a significant contribution to the state of the art as a successful green synthesis methodology utilizing Bromelia pinguin L. and contribute directly to an alternative for the removal of organic contaminants in water on an industrial scale. Moreover, the use of green synthesis methods promotes an environmentally sustainable solution by reducing reliance on toxic chemical reagents. These strategies help minimize ecological impact, aligning with sustainability goals and the preservation of water resources. The application of ZnO nanoparticles synthesized through green methods transcends environmental remediation, generating broader social impacts, including the enhancement of urban habitability. By addressing water pollution, this approach promotes cleaner water resources, which are crucial for sustaining urban life and aligning with global sustainability goals.

Author Contributions

M.d.J.C.-C.: Conceptualization, Writing—Original Draft, Supervision, and Investigation. H.E.G.G.: Methodology, Software and Data Curation. V.M.O.C.: Resources, Formal analysis and Investigation. H.G.F.: Conceptualization and Writing—Review and Editing. J.B.L.M.: Methodology and Validation. M.L.M.: Visualization and Project administration. M.O.C.: Methodology, Software, and Data Curation. P.A.L.M.: Investigation, Supervision, Writing—Review and Editing, and Visualization. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author(s).

Acknowledgments

The authors thank the support of the PIFIP 2024 Project granted by the Universidad Autónoma de Occidente (UAdeO), Unidad Regional Los Mochis. A la Convocatoria de Movilidad Academica 2024 de la Universidad Autonoma de Baja California.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Haseena, M.; Malik, M.F.; Javed, A.; Arshad, S.; Asif, N.; Zulfiqar, S.; Hanif, J. Water Pollution and Human Health. Environ. Risk Assess. Remediat. 2017, 1. [Google Scholar] [CrossRef]
  2. Leyva Morales, J.B.; Valdez Torres, J.B.; Bastidas Bastidas, P.d.J.; Angulo Escalante, M.Á.; Sarmiento Sánchez, J.I.; Barraza Lobo, A.L.; Olmeda Rubio, C.; Chaidez Quiroz, C. Monitoring of Pesticides Residues in Northwestern Mexico Rivers. Acta Univ. 2017, 27, 45–54. [Google Scholar] [CrossRef]
  3. Schweitzer, L.; Noblet, J. Water Contamination and Pollution. In Green Chemistry; Elsevier: Amsterdam, The Netherlands, 2018; pp. 261–290. [Google Scholar]
  4. Sánchez, M.B.; Uribe, C. Contaminación de Los Ambientes Acuáticos Generados Por La Industria Textil. Rev. Campus 2018, 23, 129–144. [Google Scholar] [CrossRef]
  5. Zaruma, P.; Proal, J.; Hernández, I.C.; Salas, H.I. Los Colorantes Textiles Industriales y Tratamientos Óptimos de Sus Efluentes de Agua Residual: Una Breve Revisión. Available online: https://publicaciones.ucuenca.edu.ec/ojs/index.php/quimica/article/view/2216 (accessed on 1 December 2024).
  6. Manzoor, J.; Sharma, M. Impact of Textile Dyes on Human Health and Environment. In Impact of Textile Dyes on Public Health and the Environment; IGI Global: Hershey, PA, USA, 2020; pp. 162–169. [Google Scholar]
  7. Srivastava, R.; Sofi, I.R. Impact of Synthetic Dyes on Human Health and Environment. In Impact of Textile Dyes on Public Health and the Environment; IGI Global: Hershey, PA, USA, 2020; pp. 146–161. [Google Scholar]
  8. Bhattacharjee, N.; Som, I.; Saha, R.; Mondal, S. A Critical Review on Novel Eco-Friendly Green Approach to Synthesize Zinc Oxide Nanoparticles for Photocatalytic Degradation of Water Pollutants. Int. J. Environ. Anal. Chem. 2022, 104, 489–516. [Google Scholar] [CrossRef]
  9. Hoseinpour, V.; Ghaemi, N. Green Synthesis of Manganese Nanoparticles: Applications and Future Perspective–A Review. J. Photochem. Photobiol. B Biol. 2018, 189, 234–243. [Google Scholar] [CrossRef]
  10. Nadeem, M.; Tungmunnithum, D.; Hano, C.; Abbasi, B.H.; Hashmi, S.S.; Ahmad, W.; Zahir, A. The Current Trends in the Green Syntheses of Titanium Oxide Nanoparticles and Their Applications. Green Chem. Lett. Rev. 2018, 11, 492–502. [Google Scholar] [CrossRef]
  11. Akintelu, S.A.; Folorunso, A.S.; Folorunso, F.A.; Oyebamiji, A.K. Green Synthesis of Copper Oxide Nanoparticles for Biomedical Application and Environmental Remediation. Heliyon 2020, 6, e04508. [Google Scholar] [CrossRef] [PubMed]
  12. Dowlath, M.J.H.; Musthafa, S.A.; Khalith, S.B.M.; Varjani, S.; Karuppannan, S.K.; Ramanujam, G.M.; Arunachalam, A.M.; Arunachalam, K.D.; Chandrasekaran, M.; Chang, S.W. Comparison of Characteristics and Biocompatibility of Green Synthesized Iron Oxide Nanoparticles with Chemical Synthesized Nanoparticles. Environ. Res. 2021, 201, 111585. [Google Scholar] [CrossRef]
  13. Ortíz, M.d.J.R.; Valencia, R.H.; Parra, G.A.; Morales, P.A.L. Síntesis Verde de Materiales Nanoestructurados de ZnO En La Degradación de Contaminantes Orgánicos Por Medio de La Fotocatálisis Heterogénea. J. Technol. Sci. 2021, 4, 299–313. [Google Scholar]
  14. Vasantharaj, S.; Sathiyavimal, S.; Senthilkumar, P.; Kalpana, V.N.; Rajalakshmi, G.; Alsehli, M.; Elfasakhany, A.; Pugazhendhi, A. Enhanced Photocatalytic Degradation of Water Pollutants Using Bio-Green Synthesis of Zinc Oxide Nanoparticles (ZnO NPs). J. Environ. Chem. Eng. 2021, 9, 105772. [Google Scholar] [CrossRef]
  15. Sayadi, M.H.; Ghollasimood, S.; Ahmadpour, N.; Homaeigohar, S. Biosynthesis of the ZnO/SnO2 Nanoparticles and Characterization of Their Photocatalytic Potential for Removal of Organic Water Pollutants. J. Photochem. Photobiol. A Chem. 2022, 425, 113662. [Google Scholar] [CrossRef]
  16. Luque-Morales, P.A.; Lopez-Peraza, A.; Nava-Olivas, O.J.; Amaya-Parra, G.; Baez-Lopez, Y.A.; Orozco-Carmona, V.M.; Garrafa-Galvez, H.E.; Chinchillas-Chinchillas, M.d.J. ZnO Semiconductor Nanoparticles and Their Application in Photocatalytic Degradation of Various Organic Dyes. Materials 2021, 14, 7537. [Google Scholar] [CrossRef] [PubMed]
  17. Nagajyothi, P.C.; Prabhakar Vattikuti, S.V.; Devarayapalli, K.C.; Yoo, K.; Shim, J.; Sreekanth, T.V.M. Green Synthesis: Photocatalytic Degradation of Textile Dyes Using Metal and Metal Oxide Nanoparticles-Latest Trends and Advancements. Crit. Rev. Environ. Sci. Technol. 2020, 50, 2617–2723. [Google Scholar] [CrossRef]
  18. Santhosh, P.B.; Genova, J.; Chamati, H. Green Synthesis of Gold Nanoparticles: An Eco-Friendly Approach. Chemistry 2022, 4, 345–369. [Google Scholar] [CrossRef]
  19. Wang, G.; Zhao, K.; Gao, C.; Wang, J.; Mei, Y.; Zheng, X.; Zhu, P. Green Synthesis of Copper Nanoparticles Using Green Coffee Bean and Their Applications for Efficient Reduction of Organic Dyes. J. Environ. Chem. Eng. 2021, 9, 105331. [Google Scholar] [CrossRef]
  20. Singh, J.; Dutta, T.; Kim, K.-H.; Rawat, M.; Samddar, P.; Kumar, P. ‘Green’ Synthesis of Metals and Their Oxide Nanoparticles: Applications for Environmental Remediation. J. Nanobiotechnol. 2018, 16, 84. [Google Scholar] [CrossRef]
  21. Hornung-Leoni, C.T. Avances Sobre Usos Etnobotánicos de Las Bromeliaceae En Latinoamérica. Bol. Latinoam. Caribe Plantas Med. Aromat. 2011, 10, 297–314. [Google Scholar]
  22. Rojas-García, A.R.; de los Ángeles Maldonado-Peralta, M. Cualidades Morfológicas de Frutos de Tres Especies de Bromelia (Bromeliaceae). Rev. Fitotec. Mex. 2021, 44, 521. [Google Scholar] [CrossRef]
  23. Abreu Payrol, J.; Miranda Martínez, M.; Toledo Carrabeo, G.; Castillo García, O. Actividad Farmacológica Preliminar Del Fruto de Bromelia pinguin L.(Piña de Ratón). Rev. Cuba. Farm. 2001, 35, 56–60. [Google Scholar]
  24. Efigenia, M.-G.; Miguel, A.-E.L.; Osiris, M.-O.A.; Pay, A.; Alberto, S.-B.J.; de Lourdes, G.-M.M. Physiological and Physicochemical Behavior of Guamara (Bromelia pinguin ) and Cocuixtle (Bromelia karatas) Fruits, as Well as the Antibacterial Effect of Their Pre-Purified Proteases. Emir. J. Food Agric. 2021, 277–286. [Google Scholar] [CrossRef]
  25. Meza-Espinoza, L.; de Lourdes García-Magaña, M.; de los Ángeles Vivar-Vera, M.; Sáyago-Ayerdi, S.G.; Chacón-López, A.; Becerra-Verdín, E.M.; Muy-Rangel, M.D.; Montalvo-González, E. Aspectos Etnobotánicos, Nutricionales y Actividad Biológica de Extractos de Frutos Del Género Bromelia. Rev. Fitotec. Mex. 2017, 40, 425–437. [Google Scholar] [CrossRef]
  26. Sadiq, H.; Sher, F.; Sehar, S.; Lima, E.C.; Zhang, S.; Iqbal, H.M.N.; Zafar, F.; Nuhanović, M. Green Synthesis of ZnO Nanoparticles from Syzygium Cumini Leaves Extract with Robust Photocatalysis Applications. J. Mol. Liq. 2021, 335, 116567. [Google Scholar] [CrossRef]
  27. Soto-Robles, C.A.; Luque, P.A.; Gómez-Gutiérrez, C.M.; Nava, O.; Vilchis-Nestor, A.R.; Lugo-Medina, E.; Ranjithkumar, R.; Castro-Beltrán, A. Study on the Effect of the Concentration of Hibiscus Sabdariffa Extract on the Green Synthesis of ZnO Nanoparticles. Results Phys. 2019, 15, 102807. [Google Scholar] [CrossRef]
  28. Sasi, S.; Fasna, P.H.F.; Sharmila, T.K.B.; Chandra, C.S.J.; Antony, J.V.; Raman, V.; Nair, A.B.; Ramanathan, H.N. Green Synthesis of ZnO Nanoparticles with Enhanced Photocatalytic and Antibacterial Activity. J. Alloys Compd. 2022, 924, 166431. [Google Scholar] [CrossRef]
  29. Sahai, A.; Goswami, N. Structural and Vibrational Properties of ZnO Nanoparticles Synthesized by the Chemical Precipitation Method. Phys. E Low-dimens. Syst. Nanostruct. 2014, 58, 130–137. [Google Scholar] [CrossRef]
  30. Sato-Berrú, R.Y.; Vázquez-Olmos, A.; Fernández-Osorio, A.L.; Sotres-Martínez, S. Micro-Raman Investigation of Transition-metal-doped ZnO Nanoparticles. J. Raman Spectrosc. An Int. J. Orig. Work all Asp. Raman Spectrosc. Incl. High. Order Process. Brillouin Rayleigh Scatt. 2007, 38, 1073–1076. [Google Scholar] [CrossRef]
  31. Karthik, K.V.; Raghu, A.V.; Reddy, K.R.; Ravishankar, R.; Sangeeta, M.; Shetti, N.P.; Reddy, C.V. Green Synthesis of Cu-Doped ZnO Nanoparticles and Its Application for the Photocatalytic Degradation of Hazardous Organic Pollutants. Chemosphere 2022, 287, 132081. [Google Scholar] [CrossRef]
  32. Pillai, A.M.; Sivasankarapillai, V.S.; Rahdar, A.; Joseph, J.; Sadeghfar, F.; Anuf A, R.; Rajesh, K.; Kyzas, G.Z. Green Synthesis and Characterization of Zinc Oxide Nanoparticles with Antibacterial and Antifungal Activity. J. Mol. Struct. 2020, 1211, 128107. [Google Scholar] [CrossRef]
  33. Thirumoorthy, G.S.; Balasubramaniam, O.; Kumaresan, P.; Muthusamy, P.; Subramani, K. Tetraselmis Indica Mediated Green Synthesis of Zinc Oxide (ZnO) Nanoparticles and Evaluating Its Antibacterial, Antioxidant, and Hemolytic Activity. Bionanoscience 2021, 11, 172–181. [Google Scholar] [CrossRef]
  34. Sagar Raut, D.P.; Thorat, R. Green Synthesis of Zinc Oxide (ZnO) Nanoparticles Using OcimumTenuiflorum Leaves. Int. J. Sci. Res. 2015, 4, 1225–1228. [Google Scholar]
  35. Duan, H.; Wang, D.; Li, Y. Green Chemistry for Nanoparticle Synthesis. Chem. Soc. Rev. 2015, 44, 5778–5792. [Google Scholar] [CrossRef]
  36. Luque, P.A.; Garrafa-Gálvez, H.E.; Nava, O.; Olivas, A.; Martínez-Rosas, M.E.; Vilchis-Nestor, A.R.; Villegas-Fuentes, A.; Chinchillas-Chinchillas, M.J. Efficient Sunlight and UV Photocatalytic Degradation of Methyl Orange, Methylene Blue and Rhodamine B, Using Citrus×paradisi Synthesized SnO2 Semiconductor Nanoparticles. Ceram. Int. 2021, 47, 23861–23874. [Google Scholar] [CrossRef]
  37. Suresh, D.; Shobharani, R.M.; Nethravathi, P.C.; Pavan Kumar, M.A.; Nagabhushana, H.; Sharma, S.C. Artocarpus Gomezianus Aided Green Synthesis of ZnO Nanoparticles: Luminescence, Photocatalytic and Antioxidant Properties. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 141, 128–134. [Google Scholar] [CrossRef]
  38. Gawade, V.V.; Gavade, N.L.; Shinde, H.M.; Babar, S.B.; Kadam, A.N.; Garadkar, K.M. Green Synthesis of ZnO Nanoparticles by Using Calotropis Procera Leaves for the Photodegradation of Methyl Orange. J. Mater. Sci. Mater. Electron. 2017, 28, 14033–14039. [Google Scholar] [CrossRef]
  39. Liu, Y.C.; Li, J.; Ahn, J.; Pu, J.; Rupa, E.J.; Huo, Y.; Yang, D.C. Biosynthesis of Zinc Oxide Nanoparticles by One-Pot Green Synthesis Using Fruit Extract of Amomum Longiligulare and Its Activity as a Photocatalyst. Optik 2020, 218, 165245. [Google Scholar] [CrossRef]
  40. Chakraborty, S.; Farida, J.J.; Simon, R.; Kasthuri, S.; Mary, N.L. Averrhoe Carrambola Fruit Extract Assisted Green Synthesis of Zno Nanoparticles for the Photodegradation of Congo Red Dye. Surf. Interfaces 2020, 19, 100488. [Google Scholar] [CrossRef]
  41. Zhao, S.; Sun, Y.; Lü, X.; Li, Q. Energy Consumption and Product Release Characteristics Evaluation of Oil Shale Non-Isothermal Pyrolysis Based on TG-DSC. J. Pet. Sci. Eng. 2020, 187, 106812. [Google Scholar] [CrossRef]
  42. Channa, I.A.; Ashfaq, J.; Gilani, S.J.; Shah, A.A.; Chandio, A.D.; Jumah, M.N. UV Blocking and Oxygen Barrier Coatings Based on Polyvinyl Alcohol and Zinc Oxide Nanoparticles for Packaging Applications. Coatings 2022, 12, 897. [Google Scholar] [CrossRef]
  43. Kumar, A. Sol Gel Synthesis of Zinc Oxide Nanoparticles and Their Application as Nano-Composite Electrode Material for Supercapacitor. J. Mol. Struct. 2020, 1220, 128654. [Google Scholar] [CrossRef]
  44. Khan, M.S.; Dhavan, P.P.; Jadhav, B.L.; Shimpi, N.G. Ultrasound-Assisted Green Synthesis of Ag-Decorated ZnO Nanoparticles UsingExcoecaria AgallochaLeaf Extract and Evaluation of Their Photocatalytic and Biological Activity. ChemistrySelect 2020, 5, 12660–12671. [Google Scholar] [CrossRef]
  45. Naiel, B.; Fawzy, M.; Halmy, M.W.A.; Mahmoud, A.E.D. Green Synthesis of Zinc Oxide Nanoparticles Using Sea Lavender (Limonium pruinosum L. Chaz.) Extract: Characterization, Evaluation of Anti-Skin Cancer, Antimicrobial and Antioxidant Potentials. Sci. Rep. 2022, 12, 20370. [Google Scholar] [CrossRef] [PubMed]
  46. Dmochowska, A.; Czajkowska, J.; Jędrzejewski, R.; Stawiński, W.; Migdał, P.; Fiedot-Toboła, M. Pectin Based Banana Peel Extract as a Stabilizing Agent in Zinc Oxide Nanoparticles Synthesis. Int. J. Biol. Macromol. 2020, 165, 1581–1592. [Google Scholar] [CrossRef] [PubMed]
  47. Chauhan, N.; Singh, V.; Kumar, S.; Kumari, M.; Sirohi, K. Preparation of Silver and Nitrogen Co-Doped Mesoporous Zinc Oxide Nanoparticles by Evaporation Induced Self Assembly Process to Study Their Photocatalytic Activity. J. Sol-Gel Sci. Technol. 2019, 90, 390–403. [Google Scholar] [CrossRef]
  48. Mustapha, S.; Tijani, J.O.; Ndamitso, M.M.; Abdulkareem, S.A.; Shuaib, D.T.; Mohammed, A.K.; Sumaila, A. The Role of Kaolin and Kaolin/ZnO Nanoadsorbents in Adsorption Studies for Tannery Wastewater Treatment. Sci. Rep. 2020, 10, 13068. [Google Scholar] [CrossRef]
  49. Bayrami, A.; Haghgooie, S.; Rahim Pouran, S.; Mohammadi Arvanag, F.; Habibi-Yangjeh, A. Synergistic Antidiabetic Activity of ZnO Nanoparticles Encompassed by Urtica Dioica Extract. Adv. Powder Technol. 2020, 31, 2110–2118. [Google Scholar] [CrossRef]
  50. Shebl, A.; Hassan, A.A.; Salama, D.M.; Abd El-Aziz, M.E.; Abd Elwahed, M.S.A. Green Synthesis of Nanofertilizers and Their Application as a Foliar for Cucurbita Pepo L. J. Nanomater. 2019, 2019, 3476347. [Google Scholar] [CrossRef]
  51. Miri, A.; Khatami, M.; Ebrahimy, O.; Sarani, M. Cytotoxic and Antifungal Studies of Biosynthesized Zinc Oxide Nanoparticles Using Extract of Prosopis Farcta Fruit. Green Chem. Lett. Rev. 2020, 13, 27–33. [Google Scholar] [CrossRef]
  52. Huang, G.; He, J.; Zhang, X.; Feng, M.; Tan, Y.; Lv, C.; Huang, H.; Jin, Z. Applications of Lambert-Beer Law in the Preparation and Performance Evaluation of Graphene Modified Asphalt. Constr. Build. Mater. 2021, 273, 121582. [Google Scholar] [CrossRef]
  53. Viezbicke, B.D.; Patel, S.; Davis, B.E.; Birnie III, D.P. Evaluation of the Tauc Method for Optical Absorption Edge Determination: ZnO Thin Films as a Model System. Phys. Status Solidi 2015, 252, 1700–1710. [Google Scholar] [CrossRef]
  54. Chattopadhyay, S.; Misra, K.P.; Agarwala, A.; Shahee, A.; Jain, S.; Halder, N.; Rao, A.; Babu, P.D.; Saran, M.; Mukhopadhyay, A.K. Dislocations and Particle Size Governed Band Gap and Ferromagnetic Ordering in Ni Doped ZnO Nanoparticles Synthesized via Co-Precipitation. Ceram. Int. 2019, 45, 23341–23354. [Google Scholar] [CrossRef]
  55. Mahdi Ismail, S.M.; Ahmed, S.M.; Abdulrahman, A.F.; AlMessiere, M.A. Characterization of Green Synthesized of ZnO Nanoparticles by Using Pinus Brutia Leaves Extracts. J. Mol. Struct. 2023, 1280, 135063. [Google Scholar] [CrossRef]
  56. Sarwar, N.; Shahzad, M.; Ghaffar, R.; Javed, K.; Munam, M.; Pervez, A.; Ghaffar, A. Fabrication of DSSC Based on Capsicum annuum and Tamarindus Indica Plant Seeds Extract as Natural Photosensitizers. Sol. Energy 2023, 257, 314–323. [Google Scholar] [CrossRef]
  57. Chan, Y.Y.; Pang, Y.L.; Lim, S.; Chong, W.C. Facile Green Synthesis of ZnO Nanoparticles Using Natural-Based Materials: Properties, Mechanism, Surface Modification and Application. J. Environ. Chem. Eng. 2021, 9, 105417. [Google Scholar] [CrossRef]
  58. Matinise, N.; Fuku, X.G.; Kaviyarasu, K.; Mayedwa, N.; Maaza, M. ZnO Nanoparticles via Moringa Oleifera Green Synthesis: Physical Properties & Mechanism of Formation. Appl. Surf. Sci. 2017, 406, 339–347. [Google Scholar] [CrossRef]
  59. Li, S.; Shen, Y.; Xie, A.; Yu, X.; Qiu, L.; Zhang, L.; Zhang, Q. Green Synthesis of Silver Nanoparticles Using Capsicum annuum L. Extract. Green Chem. 2007, 9, 852–858. [Google Scholar] [CrossRef]
  60. Bandeira, M.; Giovanela, M.; Roesch-Ely, M.; Devine, D.M.; da Silva Crespo, J. Green Synthesis of Zinc Oxide Nanoparticles: A Review of the Synthesis Methodology and Mechanism of Formation. Sustain. Chem. Pharm. 2020, 15, 100223. [Google Scholar] [CrossRef]
  61. Ahmed, S.; Annu; Chaudhry, S.A.; Ikram, S. A Review on Biogenic Synthesis of ZnO Nanoparticles Using Plant Extracts and Microbes: A Prospect towards Green Chemistry. J. Photochem. Photobiol. B Biol. 2017, 166, 272–284. [Google Scholar] [CrossRef]
  62. Hussein, B.Y.; Mohammed, A.M. Green Synthesis of ZnO Nanoparticles in Grape Extract: Their Application as Anti-Cancer and Anti-Bacterial. Mater. Today Proc. 2021, 42, A18–A26. [Google Scholar] [CrossRef]
  63. Anju Chanu, L.; Joychandra Singh, W.; Jugeshwar Singh, K.; Nomita Devi, K. Effect of Operational Parameters on the Photocatalytic Degradation of Methylene Blue Dye Solution Using Manganese Doped ZnO Nanoparticles. Results Phys. 2019, 12, 1230–1237. [Google Scholar] [CrossRef]
  64. Mahmoud, Z.H. Photodegradation of Methylene Blue Solution via Au Doped TiO2 Nanocomposite Catalysts Prepared Using Novel Photolysis Method. Iran. J. Chem. Chem. Eng. 2019, 38, 29–35. [Google Scholar]
  65. Han, C.; Cheng, C.; Liu, F.; Li, X.; Wang, G.; Li, J. Preparation of CdS–Ag2S Nanocomposites by Ultrasound-Assisted UV Photolysis Treatment and Its Visible Light Photocatalysis Activity. Nanotechnol. Rev. 2023, 12, 20220503. [Google Scholar] [CrossRef]
  66. Alshamsi, H.A.; Jaffer, A.A. New Hibiscus Sabdariffa L Petals Extract Based Green Synthesis of Zinc Oxide Nanoparticles for Photocatalytic Degradation of Rhodamine B Dye under Solar Light. AIP Conf. Proc. 2022, 2394, 40017. [Google Scholar] [CrossRef]
  67. Soto-Robles, C.A.; Nava, O.; Cornejo, L.; Lugo-Medina, E.; Vilchis-Nestor, A.R.; Castro-Beltrán, A.; Luque, P.A. Biosynthesis, Characterization and Photocatalytic Activity of ZnO Nanoparticles Using Extracts of Justicia Spicigera for the Degradation of Methylene Blue. J. Mol. Struct. 2021, 1225, 129101. [Google Scholar] [CrossRef]
  68. Nguyen-Hong, Y.; Luu, T.V.H.; Doan, V. Green Synthesis of Ce-doped ZnO Nanoparticles Using Hedyotis Capitellata Leaf Extract for Efficient Photocatalytic Degradation of Methyl Orange. Vietnam J. Chem. 2021, 59, 648–659. [Google Scholar] [CrossRef]
  69. Ahmad, M.; Rehman, W.; Khan, M.M.; Qureshi, M.T.; Gul, A.; Haq, S.; Ullah, R.; Rab, A.; Menaa, F. Phytogenic Fabrication of ZnO and Gold Decorated ZnO Nanoparticles for Photocatalytic Degradation of Rhodamine B. J. Environ. Chem. Eng. 2021, 9, 104725. [Google Scholar] [CrossRef]
  70. Davis, K.; Yarbrough, R.; Froeschle, M.; White, J.; Rathnayake, H. Band Gap Engineered Zinc Oxide Nanostructures via a Sol–Gel Synthesis of Solvent Driven Shape-Controlled Crystal Growth. RSC Adv. 2019, 9, 14638–14648. [Google Scholar] [CrossRef]
  71. Villarreal, R.C.; Luque-Morales, M.; Chinchillas-Chinchillas, M.J.; Luque, P.A. Langmuir-Hinshelwood-Hougen-Watson Model for the Study of Photodegradation Properties of Zinc Oxide Semiconductor Nanoparticles Synthetized by Peumus boldus. Results Phys. 2022, 36, 105421. [Google Scholar] [CrossRef]
  72. Escobar-Alarcón, L.; Solís-Casados, D.A. Desarrollo de Fotocatalizadores Basados En TiO2 En Forma de Película Delgada Para La Degradación de Moléculas Orgánicas En Solución Acuosa. Mundo Nano. Rev. Interdiscip. Nanocienc. Nanotecnol. 2021, 14, 1e–23e. [Google Scholar] [CrossRef]
Figure 1. ATR-IR analysis of the nanoparticles biosynthesized with Bromelia pinguin L.
Figure 1. ATR-IR analysis of the nanoparticles biosynthesized with Bromelia pinguin L.
Sustainability 16 10745 g001
Figure 2. Morphological analysis of the nanoparticles biosynthesized with Bromelia pinguin L. (a,b) Micrographs of BP 1%-ZnO, (c) size distribution of BP 1%-ZnO, (d,e) Micrographs of BP 2%-ZnO, (f) size distribution of BP 2%-ZnO, (g,h) Micrographs of BP 4%-ZnO and (i) size distribution of BP 4%-ZnO.
Figure 2. Morphological analysis of the nanoparticles biosynthesized with Bromelia pinguin L. (a,b) Micrographs of BP 1%-ZnO, (c) size distribution of BP 1%-ZnO, (d,e) Micrographs of BP 2%-ZnO, (f) size distribution of BP 2%-ZnO, (g,h) Micrographs of BP 4%-ZnO and (i) size distribution of BP 4%-ZnO.
Sustainability 16 10745 g002
Figure 3. XRD spectra of the nanoparticles biosynthesized with Bromelia pinguin L. (a) BP 1%-ZnO nanoparticles, (b) BP 2%-ZnO nanoparticles and (c) BP 4%-ZnO nanoparticles.
Figure 3. XRD spectra of the nanoparticles biosynthesized with Bromelia pinguin L. (a) BP 1%-ZnO nanoparticles, (b) BP 2%-ZnO nanoparticles and (c) BP 4%-ZnO nanoparticles.
Sustainability 16 10745 g003
Figure 4. TGA/DSC results of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
Figure 4. TGA/DSC results of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
Sustainability 16 10745 g004
Figure 5. BET analysis of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
Figure 5. BET analysis of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
Sustainability 16 10745 g005
Figure 6. UV–Vis spectra of nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
Figure 6. UV–Vis spectra of nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
Sustainability 16 10745 g006
Figure 7. Band gaps of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L. calculated with the help of the TAUC model.
Figure 7. Band gaps of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L. calculated with the help of the TAUC model.
Sustainability 16 10745 g007
Figure 8. Formation mechanism of nanoparticles biosynthesized with Bromelia pinguin L.
Figure 8. Formation mechanism of nanoparticles biosynthesized with Bromelia pinguin L.
Sustainability 16 10745 g008
Figure 9. Photocatalytic activity of nanoparticles biosynthesized with Bromelia pinguin L. (a) Degradation of MB under solar radiation; (b) degradation of MB under UV radiation; (c) MO degradation under solar radiation; (d) degradation of MO under UV radiation; (e) degradation of RhB under solar radiation; and (f) degradation of RhB under UV radiation.
Figure 9. Photocatalytic activity of nanoparticles biosynthesized with Bromelia pinguin L. (a) Degradation of MB under solar radiation; (b) degradation of MB under UV radiation; (c) MO degradation under solar radiation; (d) degradation of MO under UV radiation; (e) degradation of RhB under solar radiation; and (f) degradation of RhB under UV radiation.
Sustainability 16 10745 g009
Figure 10. Proposed photocatalytic degradation mechanism of ZnO nanoparticles biosynthesized with Bromelia pinguin L.
Figure 10. Proposed photocatalytic degradation mechanism of ZnO nanoparticles biosynthesized with Bromelia pinguin L.
Sustainability 16 10745 g010
Table 1. Thermogravimetric results of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
Table 1. Thermogravimetric results of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
SamplesSectionWeight lossTotal Loss
BP 1%-ZnO11.1%2.6%
20.5%
31%
BP 2%-ZnO13.2%5.7%
21.1%
31.4%
BP 4%-ZnO16%8.9%
21.6%
31.3%
Table 2. Results of the BET analysis of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
Table 2. Results of the BET analysis of the nanoparticles synthesized using 1%, 2%, and 4% of Bromelia pinguin L.
SampleSurface Area (m2/g)Pore Volume (cc/g)Pore Size (Å)
BP 1%-ZnO7.2990.0328.925
BP 2%-ZnO20.2850.0998.020
BP 4%-ZnO21.9800.1387.854
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chinchillas-Chinchillas, M.d.J.; Galvez, H.E.G.; Carmona, V.M.O.; Galindo Flores, H.; Morales, J.B.L.; Luque Morales, M.; Camacho, M.O.; Luque Morales, P.A. Sustainable Green Synthesis of ZnO Nanoparticles from Bromelia pinguin L.: Photocatalytic Properties and Their Contribution to Urban Habitability. Sustainability 2024, 16, 10745. https://doi.org/10.3390/su162310745

AMA Style

Chinchillas-Chinchillas MdJ, Galvez HEG, Carmona VMO, Galindo Flores H, Morales JBL, Luque Morales M, Camacho MO, Luque Morales PA. Sustainable Green Synthesis of ZnO Nanoparticles from Bromelia pinguin L.: Photocatalytic Properties and Their Contribution to Urban Habitability. Sustainability. 2024; 16(23):10745. https://doi.org/10.3390/su162310745

Chicago/Turabian Style

Chinchillas-Chinchillas, Manuel de Jesus, Horacio Edgardo Garrafa Galvez, Victor Manuel Orozco Carmona, Hugo Galindo Flores, Jose Belisario Leyva Morales, Mizael Luque Morales, Mariel Organista Camacho, and Priscy Alfredo Luque Morales. 2024. "Sustainable Green Synthesis of ZnO Nanoparticles from Bromelia pinguin L.: Photocatalytic Properties and Their Contribution to Urban Habitability" Sustainability 16, no. 23: 10745. https://doi.org/10.3390/su162310745

APA Style

Chinchillas-Chinchillas, M. d. J., Galvez, H. E. G., Carmona, V. M. O., Galindo Flores, H., Morales, J. B. L., Luque Morales, M., Camacho, M. O., & Luque Morales, P. A. (2024). Sustainable Green Synthesis of ZnO Nanoparticles from Bromelia pinguin L.: Photocatalytic Properties and Their Contribution to Urban Habitability. Sustainability, 16(23), 10745. https://doi.org/10.3390/su162310745

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop