image
image
user-login
Patent search/

MECHANICALLY TRIGGERED ON-DEMAND EASY-TO-FABRICATE DONOR-ACCEPTOR NANOCOMPOSITE PHOTOCATALYST

search

Patent Search in India

  • tick

    Extensive patent search conducted by a registered patent agent

  • tick

    Patent search done by experts in under 48hrs

₹999

₹399

Talk to expert

MECHANICALLY TRIGGERED ON-DEMAND EASY-TO-FABRICATE DONOR-ACCEPTOR NANOCOMPOSITE PHOTOCATALYST

ORDINARY APPLICATION

Published

date

Filed on 9 November 2024

Abstract

Mechanically triggered polymeric nanocomposites that can be manufactured on demand have the potential to enable chemical recycling, and minimize environmental pollution. It can be beneficial in tissue engineering by emulating natural photosynthesis, which depends on demand. Biodegradable polymeric material such as polylactic acid (P) possesses excellent manufacturing qualities compared to other biopolymers due to excellent tensile and flexural properties. This study aimed to fabricate an integrated donor-acceptor P-based nanocomposite flexible artificial leaf as a film photocatalyst with ultra-high tensile strength, bending strength, impact strength, and surface hardness via the blown film method. Photocatalytic fixation, an improved reductive process, has been recognized as a cost-effective and reusable technology for indoor and outdoor environmental management. Therefore, newly designed and fabrication of cost-effective integrated donor-acceptor nanocomposite flexible artificial leaf (PGT) as film photocatalyst which is the incorporation of magnesium tetra-phenyl-porphyrin (T)/aloe-vera-derived graphene (0.5, 1 and 1.5% G) into polylactic acid (P). Among these, 1% PGT photocatalyst-developed integrated donor-acceptor nanocomposite artificial leaf as a film photocatalyst extremely well, producing significant levels of solar light active 1,4-NADH regeneration (61.09 ± 0.59%) and its consumption by the formate dehydrogenase enzyme in the sole creation of formic acid (HCOOH~146.62 ± 1.6 µmol) from CO2. Because of its extremely high tensile strength (25.322MPa), tensile load (589.49 Newton), strain (11.755%), bending strength (32.244MPa), and impact energy (2.4615 Joule), PGT nanocomposite serve as a suitable material for tissue implants for various applications. Therefore, 1% PGT nanocomposite flexible artificial leaf as a film photocatalyst has a remarkable ability to fix CO2 into HCOOH when compared to other flexible artificial leaf as a film photocatalyst (0.5 and 1.5% PGT). Overall, the outcome demonstrates the potential and adaptability of these P-based nanocomposite artificial leaves (PGT), emphasizing their importance in photocatalysis, solar chemical synthesis, and scaffold-based tissue engineering.

Patent Information

Application ID202411086442
Invention FieldBIO-CHEMISTRY
Date of Application09/11/2024
Publication Number47/2024

Inventors

NameAddressCountryNationality
DR. RAJESH KUMAR YADAVDEPARTMENT OF CHEMISTRY AND ENVIRONMENTAL SCIENCE, MADAN MOHAN MALAVIYA UNIVERSITY OF TECHNOLOGY, GORAKHPUR-273010, INDIAIndiaIndia
DR. RAJESH KUMAR VERMADEPARTMENT OF MECHANICAL ENGINEERING, HARCOURT BUTLER TECHNICAL UNIVERSITY KANPURIndiaIndia
DR. KULDEEP KUMARDEPARTMENT OF CHEMISTRY AND ENVIRONMENTAL SCIENCE, MADAN MOHAN MALAVIYA UNIVERSITY OF TECHNOLOGY, GORAKHPUR-273010, INDIAIndiaIndia
DR. VANDNA RANI VERMADEPARTMENT OF COMPUTER SCIENCE AND ENGINEERING, GALGOTIAS COLLEGE OF ENGINEERING AND TECHNOLOGY, GREATER NOIDA, U. P., INDIAIndiaIndia

Applicants

NameAddressCountryNationality
MADAN MOHAN MALAVIYA UNIVERSITY OF TECHNOLOGYMADAN MOHAN MALAVIYA UNIVERSITY OF TECHNOLOGY, GORAKHPUR UTTAR PRADESHIndiaIndia

Specification

DESC:FIELD OF THE INVENTION
The present invention relates to mechanically triggered on-demand easy-to-fabricate donor-acceptor nanocomposite photocatalyst.
BACKGROUND OF THE INVENTION
References which are cited in the present disclosure are not necessarily prior art and therefore their citation does not constitute an admission that such references are prior art in any jurisdiction. All publications, patents and patent applications herein are incorporated by reference to the same extent as if each individual or patent application was specifically and individually indicated to be incorporated by reference.
The 400-500 mL of air are inspired and exhaled by adults during each respiratory cycle [1]. In the majority, 90% of a person's time is spent indoors, in homes, businesses, malls, gyms, and educational institutions [2]. As a result, the quality of life and general health can be considerably impacted by indoor air quality [3]. Human health is severely impacted by poor indoor air quality, especially when it comes to indoor air microbiomes, such as those containing pathogenic bacteria [4] and severe acute respiratory syndrome coronavirus [5-7]. During the COVID-19 pandemic, individuals were predominantly confined to their homes due to harmful air pathogens. Increase in air pollutants not only disrupt the environmental balance but also promote the growth of harmful pathogens in the air. These pollutants comprise a variety of greenhouse gases, including methane, carbon dioxide, nitrous oxide, perfluorocarbons, sulfur hexafluoride, nitrogen trifluoride, and many more8. Among all pollutants, anthropogenic CO2 accounts for the main indoor and outdoor environmental problems such as global warming [8], carcinogenicity [9], and adverse health effects [10]. One of the major issues that has surfaced in recent years is global warming, largely brought about by an increase in atmospheric level of carbon dioxide (CO2) [1-4]. Every day, a significant amount of CO2 is emitted into the atmosphere through a variety of processes, notably by the burning of fossil fuels. Reliance on fossil fuels disrupt the carbon cycle and increases the excessive consumption of natural resources [5,11,12].
Through a process known as photosynthesis, plants turn CO2 into sugar to keep the carbon cycle going, but overuse of fossil fuels, industrialization, deforestation, and urbanization all have a huge impact on this process. A rise in global temperature of 1.9 °C is predicted to occur by 2100 if anthropogenic CO2 levels reach approximately 590 ppm [13]. The existence of all life on Earth may be seriously impacted by the fact that the temperature rising in the polar areas would be three times higher than in other locations [5]. Most nations have been developing sustainable and renewable energy generation technologies in significant quantities in response to this[14]. In recent times, the scientific community has placed great emphasis on the extraction and utilization of CO2 [8-10,15-17]. Numerous research initiatives have been set up to collect and store surplus CO2 [18-21]. The primary practical limitations of these techniques are the potential for CO2 leakage, high energy requirements, and intricate designs [5]. Therefore, recycling or converting CO2 into fuels and other goods with added value is a desirable way to alleviate the energy crisis and global warming without preventing urbanization and growth.
Although the conversion of CO2 is a difficult scientific endeavour, there are many advantages. CO2 conversion has previously been achieved using flexible artificial leaf photocatalysis [22-26], thermocatalysis [27-31], radiolysis [32-34], and biochemical [35,36]. Among these, photocatalysis is shown to be the most effective way to use solar light irradiation to renovate CO2 into specific gaseous and liquid products, particularly for liquid products at ambient temperature and pressure [37-39]. Because this technology replicates the energy cycle of nature, it is also known as artificial photosynthesis [40]. In addition to using sustainable solar energy, flexible artificial leaf photocatalysis (FALPC) offers other important advantages, including control over product selectivity, environmental compatibility, and economic viability [41-43]. The conversion reaction of CO2, a linear, chemically highly stable molecule with low electron affinity, is dictated by nucleophilic assaults on the carbon atom [44]. More than 750 kJ/mol is required for the energy dissociation to break the C=O bond [45]. Since the process is thermodynamically an uphill one, more energy must be added in order to disrupt the C=O bond, overcome resistance loss, and create band bending [46].
Natural photosynthesis uses energy absorbed from solar light by light harvestor/photosensitizers to accomplish uphill reaction processes, which is similar to our designed artificial photosynthetic route due to negative free energy. To convert CO2, photogenerated electrons and holes must have greater and lower energy levels than the overpotential of H2O/O2 (0.82 at neutral pH) and CO2/HCOOH (-0.61 V at neutral pH) [37,47,48]. Photocatalysts having a broad energy gap 2.88 eV are required to execute redox processes such as sacrificial agent oxidation and CO2 fixation [49]. The product amount of CO2 transformation is determined by the various reported photocatalytic materials [50], which have weaker photocatalytic characteristics [51].
This paper details the successful creation of an ultra-high tensile, bending, impact strength, surface hardness, and highly capable light-harvesting nanocomposites flexible artificial leaf photocatalyst for CO2 fixation using solar energy. The high tensile strength (25.322MPa), tensile load (589.49 Newton), strain (11.755%), bending strength (32.244MPa), and impact energy (2.4615 Joule) of PGT nanocomposite aid the material to show promise for potential applications in tissue implants. This study used 0.5, 1, and 1.5% PGT nanocomposites to create flexible artificial leaves. In addition to simple and cost-effective preparation, the 'T' system has an efficient energy transfer ability, high molar extinction coefficient, a long-excited state lifetime, and photochemical stability. The resultant PGT nanocomposites flexible artificial leaf photocatalyst obtained by blown film method exhibited superior performance due to ultra-high tensile and bending strength over the 0.5, and 1.5 % PGT nanocomposites flexible artificial leaf photocatalysts in regeneration of 1,4-NADH regeneration, and exclusive formation of formic acid (HCOOH) from CO2, enzymatically, using solar light [52-54]. The schematic representation of CO2 conversion using inexhaustible solar energy is depicted in scheme 1[55].
Several patents issued for photocatalysts but none of these are related to the present invention. Patent US11518689B2 relates to a method, comprising: removing a per- or polyfluoroalkyl substance (PFAS) from a sample by combining the sample with a composite sorbent, thereby sorbing the PFAS to the sorbent to form a PFAS-sorbent, the composite sorbent comprising at least two different materials selected from (a) a metal-organic framework (MOF), a covalent organic framework (COF), a covalent organic polymer (COP), zeolites, mesoporous silica, hierarchical porous carbon in combination with (b) at least one of a polymer, a zeolite, a covalent organic framework, mesoporous silica, a hierarchical porous carbon, a photocatalyst, a carbon nanotube, graphite, graphene, graphene oxide, a Prussian blue analog, or a metal oxide; and separating the PFAS-sorbent from the sample, wherein (i) the MOF, if present, does not comprise [Zr6O4(OH)4] and 1,4-benzodicarboxylic acid (UiO-66), and/or (ii) the polymer, if present, is not poly(ethylene-co-vinyl acetate).
Another patent RU2706318C2 relates to chemical industry and nanotechnology. Composite material with primary particle size of 0.1-100 mcm contains graphene oxide and 0.1-50 wt. % of iron compound retained thereon, for example Fe3O4, Fe2O3 or mixtures thereof. Size of particles of iron compound is 0.1-10 nm. In the infrared spectrum of said composite material there is virtually no absorption occurring from the OH group, absorption occurring from the C = O group, and absorption of about 701 cm-1, occurring from the Fe-O group, but there is absorption occurring from the CO group. To obtain said composite material, the corresponding raw materials are suspended in an inert solvent and the obtained suspension is irradiated with UV and visible radiation with wavelength of 100-800 nm from 1 minute to 24 hours. As a raw material of the iron compound, at least one of iron and an inorganic acid, an iron salt and a carboxylic acid, an iron salt and sulphonic acid, iron hydroxide, phenol iron, double iron salts and iron complexes. Obtained composite material is used as a photocatalyst or equipment when producing hydrogen from water and/or alcohol by irradiating with sunlight or LED white light.
Another patent US10449530B2 relates to photocatalysts for reduction of carbon dioxide and water are provided that can be tuned to produce certain reaction products, including hydrogen, alcohol, aldehyde, and/or hydrocarbon products. These photocatalysts can form artificial photosystems and can be incorporated into devices that reduce carbon dioxide and water for production of various fuels. Doped wide-bandgap semiconductor nanotubes are provided along with synthesis methods. A variety of optical, electronic and magnetic dopants (substitutional and interstitial, energetically shallow and deep) are incorporated into hollow nanotubes, ranging from a few dopants to heavily-doped semiconductors. The resulting wide-bandgap nanotubes, with desired electronic (p- or n-doped), optical (ultraviolet bandgap to infrared absorption in co-doped nanotubes), and magnetic (from paramagnetic to ferromagnetic) properties, can be used in photovoltaics, display technologies, photocatalysis, and spintronic applications.
Another patent US9205420B2 relates to nanostructures and compositions comprising nanostructures, methods of making and using the nanostructures, and related systems. In some embodiments, a nanostructure comprises a first region and a second region, wherein a first photocatalytic reaction (e.g., an oxidation reaction) can be carried out at the first region and a second photocatalytic reaction (e.g., a reduction reaction) can be carried out at the second region. In some cases, the first photocatalytic reaction is the formation of oxygen gas from water and the second photocatalytic reaction is the formation of hydrogen gas from water. In some embodiments, a nanostructure comprises at least one semiconductor material, and, in some cases, at least one catalytic material and/or at least one photosensitizing agent.
Another patent US7541509B2 discloses a photocatalyst nanocomposite which can be used to destroying biological agents includes a carbon nanotube core, and a photocatalyst coating layer covalently or ionically bound to a surface of the nanotube core. The coating layer has a nanoscale thickness. A method of forming photocatalytic nanocomposites includes the steps of providing a plurality of dispersed carbon nanotubes, chemically oxidizing the nanotubes under conditions to produce surface functionalized nanotubes to provide C and O including groups thereon which form ionic or covalent bonds to metal oxides, and processing a metal oxide photocatalyst sol-gel precursor in the presence of the nanotubes, wherein a nanoscale metal oxide photocatalyst layer becomes covalently or ionically bound to the nanotubes.
Another patent CN107349937B discloses a preparation method of a graphene-based bimetallic sulfide nano composite photocatalyst, which is characterized in that a series of graphene-based bimetallic sulfide composite photocatalysts are synthesized by a simple and mild one-step hydrothermal method, and the morphology size and the oriented growth on the surface of graphene of bimetallic sulfide are regulated and controlled by controlling the hydrothermal reaction temperature, the reaction time, the addition amount of graphene and the content of a metal salt compound in a composite system; in addition, the preparation method provided by the invention is simple in process operation, low in raw material price and suitable for large-scale industrial production.
Another patent CN102921416B relates to a silver doped grapheme-zinc oxide nano composite photocatalytic material and a method for preparing the same and belongs to the technical field of nano composite materials and photocatalysis. Graphite oxide is subjected to ultrasonic dispersion to obtain a dispersion solution of graphene oxide; precursors of silver ions and zinc ions are added in the dispersion solution of the graphene oxide, and the mixture is placed into a reaction kettle to be subjected to hydro-thermal treatment to be prepared into the silver doped grapheme-zinc oxide nano composite photocatalytic material in situ after the pH is regulated to an alkaline condition. Photocatalytic degradation experiments show that the silver doped grapheme-zinc oxide nano composite photocatalytic material prepared through the method has good adsorption and visible light photocatalytic degradation effects on rhodamine B and is an ideal nano composite photocatalytic material.
OBJECTS OF THE INVENTION
Main object of the present invention is to develop mechanically triggered polymeric nanocomposites that can be manufactured on demand.
Another object of the present invention is to enable chemical recycling to minimize environmental pollution.
Another object of the present invention is to utilize biodegradable polylactic acid (P) as the primary polymer due to its superior manufacturing properties.
Another object of the present invention is to fabricate an integrated donor-acceptor P-based nanocomposite flexible artificial leaf as a film photocatalyst.
Another object of the present invention is to employ photocatalytic fixation to enhance CO2 conversion processes as a cost-effective, reusable environmental management solution.
SUMMARY OF THE INVENTION
This summary is provided to introduce a selection of concepts, in a simplified format, that are further described in the detailed description of the invention.
This summary is neither intended to identify key or essential inventive concepts of the invention and nor is it intended for determining the scope of the invention.
To further clarify advantages and features of the present invention, a more particular description of the invention will be rendered by reference to specific embodiments thereof, which is illustrated in the appended drawings.
During the fabrication processes to prevent agglomeration, G powder was first dispersed in chloroform (CHCl3) and then sonicated at 45 kHz for an hour. P pellets were then added to the mixture, along with CHCl3, and the synthesized G powder was stirred at 170°C for two hours and 30 minutes. The resin blend was poured into a mold sized 100 × 100 × 3 mm and dried for 24 hours under solar light. The fabricated composite was cut for mechanical strength testing in accordance with ASTM standards, followed by tensile, flexural, impact, and hardness tests, as well as 1,4-NADH regeneration. For comparative mechanical strength testing, four types of specimens were developed: P neat, 0.5, 1, and 1.5% PGT nanocomposite flexible artificial leaf photocatalyst.
Herein enclosed a mechanically triggered on-demand easy-to-fabricate donor-acceptor nanocomposite photocatalyst comprising of
Nicotinamide adenine dinucleotide (ß-NAD+), formate dehydrogenase enzyme, 2,2'-bipyridine, (pentethylcyclopentadienyl) rhodium (III) chloride dimer, N, N-dimethylformamide, polylactic acid, chloroform, cyclohexane, silica gel, and magnesium chloride (MgCl2), iodine, pyrrole, dichloromethane, p-chloranil, benzaldehyde, ascorbic acid, monobasic and dibasic phosphate buffer.
A method for the preparation of mechanically triggered on-demand easy-to-fabricate donor-acceptor nanocomposite photocatalyst comprising the steps of
cutting raw aloe vera into small pieces and washed with DI water several times after the green skin was carefully removed;
grounding the internal part of the aloe vera leaf in the mixer along with ethanol;
stirring the prepared mixture at 550 rpm along with 150°C continues until the mixed solution becomes dry;
transferring a specific amount of aloe vera gel to a crucible and heated in a muffle furnace at 500°C for four hours after drying;
removing the crucible from the muffle furnace and allowed to cool at room temperature after the heating process;
obtaining product aloe vera drive graphene;
adding benzaldehyde (1 mmol), iodine (0.21 equiv), and pyrrole (1 mmol) in turn to 10 mL of dichloromethane (DCM);
after adding the para-chloranil (0.75 equivalent), a second activation (300 W, 35 0C) was carried out;
using DCM/petroleum ether (1:1) as the eluent, the mixture was purified by flash chromatography;
obtaining a purple solid (59 mg, 36%) (tetra-phenyl-porphyrin);
allowing the solution containing 0.25 g of tetra-phenyl-porphyrin and 0.28 g of MgCl2 in 10 mL of DMF to reflux for about 1 hour;
allowing the solution to cool after refluxing;
adding 200 mL of distilled water to the cooled solution and it was maintained in an ice bath for 15 minutes;
following filtering, the solid precipitate of Mg-TPP was obtained; and
rinsing the product with distilled water and left to dry;
dispersing Graphene powder in chloroform (CHCl3) and then sonicated at 45 kHz for an hour;
adding PLA pellets along with CHCl3 and Mg-TPP to the mixture;
stirring the mixture at 170°C for two hours and 30 minutes;
pouring the resin blend into a mould sized 100 × 100 × 3 mm and dried for 24 hours under solar light;
cutting the fabricated composite for mechanical strength testing in accordance with ASTM standards.
The testing mechanical strength in accordance with ASTM standards, followed by tensile, flexural, impact, and hardness tests, as well as 1,4-NADH regeneration.
For comparative mechanical strength testing, four types of specimens were developed: P neat, 0.5, 1, and 1.5% PGT nanocomposite flexible artificial leaf photocatalyst.
The production of formic acid from CO2 was also carried out in a quartz reactor at ambient temperature in an inert environment utilising a solar light source.
Using a solar light source and an inert atmosphere, photochemical regeneration of 1,4-NADH was carried out at ambient temperature in a quartz reactor.
BRIEF DESCRIPTION OF THE DRAWINGS
The illustrated embodiments of the subject matter will be understood by reference to the drawings, wherein like parts are designated by like numerals throughout. The following description is intended only by way of example, and simply illustrates certain selected embodiments of devices, systems, and methods that are consistent with the subject matter as claimed herein, wherein:
Figure 1. Synthesis of Aloe vera derived graphene (G) from green aloe-vera leaf.
Figure 2. Synthesis of Magnesium tetraphenylporphyrin (Mg-TPP~T) via condensation method.
Figure 3. Development processes PGT nanocomposite flexible artificial leaf photocatalyst.
Figure 4. (a) Diffuse reflectance spectra of G, T, and 1%PGT photocatalysts, and (b) Optical band gap calculation of T, and 1% PGT light harvesting materials by Kubelka-Munk (K-M) method.
Figure 5. (a) FTIR spectra of Aloe vera drive G, P Neat, P@G, P@T, 0.5, 1, 1.5 % PGT (b) X-ray diffraction of G, P neat, P@G, P@T, and 1% PGT nanocomposite flexible artificial leaf photocatalyst, and (c) TGA plot comparing the mass loss of P neat, P@G and 1% PGT nanocomposite photocatalyst.
Figure 6. (a) FTIR spectra of Aloe vera drive G, P Neat, P@G, P@T, 0.5, 1, 1.5 % PGT (b) X-ray diffraction of G, P neat, P@G, P@T, and 1% PGT nanocomposite flexible artificial leaf photocatalyst, and (c) TGA plot comparing the mass loss of P neat, P@G and 1% PGT nanocomposite photocatalyst.
Figure 7. SEM analysis of (a) P Neat (b) P@G (c) P@T (d) 1% PGT nanocomposite flexible artificial leaf photocatalyst.
Figure 8. Universal testing machine setup (UTM Model- EZ50) for tensile and flexural test.
Figure 9. Universal testing machine setup (UTM Model- EZ50) for tensile and flexural test.
Figure 10. Stress vs Strain curve of tensile test for nanocomposite flexible artificial leaf photocatalyst (a) 0 % PGT (b) 0.5 % PGT (c) 1 % PGT (d) 1.5 % PGT nanocomposites flexible film as an artificial leaf photocatalyst.
Figure 11. Flexural testing specimens (a) Prepared specimen (b) Broken specimen.
Figure 12. Flexural Stress vs Strain curve of flexural test for nanocomposite flexible artificial leaf photocatalyst (a) 0 % PGT (b) 0.5 % PGT (c) 1 % PGT (d) 1.5 % PGT.
Figure 13. (a) Impact test setup (b) Impact strength of a tested specimen concerning varying wt.% of G.
Figure 14. The hardness of P neat, 0.5 %, 1 %, and 1.5 PGT nanocomposite photocatalyst.
Figure 15. The photocatalytic activity of P, T, 0.5% PGT, 1% PGT, and 1.5% PGT in solar light-driven artificial photosynthesis of formic acid from CO2 and 1,4-NADH photoregeneration. (a) Photo regeneration of 1,4-NADH, and (b) Artificial photosynthesis of formic acid from CO2 driven by solar light.
BRIEF DESCRIPTION OF THE TABLES
Table 1. Mechanical properties of developed nanocomposite flexible artificial leaf photocatalyst.
Table. 2 Flexural test results of developed nanocomposite photocatalyst
DETAILED DESCRIPTION OF THE INVENTION
In some embodiments of the present invention, during the fabrication processes to prevent agglomeration, G powder was first dispersed in chloroform (CHCl3) and then sonicated at 45 kHz for an hour.
In some embodiments of the present invention, P pellets were then added to the mixture, along with CHCl3, and the synthesized G powder was stirred at 170°C for two hours and 30 minutes. The resin blend was poured into a mold sized 100 × 100 × 3 mm and dried for 24 hours under solar light.
In some embodiments of the present invention, the fabricated composite was cut for mechanical strength testing in accordance with ASTM standards, followed by tensile, flexural, impact, and hardness tests, as well as 1,4-NADH regeneration. For comparative mechanical strength testing, four types of specimens were developed: P neat, 0.5, 1, and 1.5% PGT nanocomposite flexible artificial leaf photocatalyst.
In some embodiments of the present invention, using a solar light source and an inert atmosphere, photochemical regeneration of 1,4-NADH was carried out at ambient temperature in a quartz reactor. Here is how the photocatalytic regeneration of 1,4-NADH was done. A quartz reactor was used to carry out the reaction.
In some embodiments of the present invention, in 3.1 mL of sodium phosphate buffer (100 mM, pH 7.0), the following components of the reaction were present: ß-NAD+ (1.24 mmol), rhodium complex Rh (0.62 mmol) synthesised as per reported paper [56], AsA (1.24 mmol), and film photocatalyst (1×1cm2). The UV-vis spectrophotometer was used to track the regeneration of 1,4-NADH.
In some embodiments of the present invention, the production of formic acid from CO2 was also carried out in a quartz reactor at ambient temperature in an inert environment utilising a solar light source. The procedure used film photocatalyst (1×1cm2), ß-NAD+ (1.24 mmol), rhodium complex Rh (0.62 mmol), and formate dehydrogenase (3 units) in 3.1 mL of sodium phosphate buffer (100 mM, pH 7.0) with AsA (1.24 mmol) in the presence of CO2.
In some embodiments of the present invention, after bubbling CO2 for 1 hour without light (light off), the reactor was exposed to visible light (light on). The quantity of formic acid was determined using UV-visible spectroscopy and HPLC.
Herein enclosed a mechanically triggered on-demand easy-to-fabricate donor-acceptor nanocomposite photocatalyst comprising of
Nicotinamide adenine dinucleotide (ß-NAD+), formate dehydrogenase enzyme, 2,2'-bipyridine, (pentethylcyclopentadienyl) rhodium (III) chloride dimer, N, N-dimethylformamide, polylactic acid, chloroform, cyclohexane, silica gel, and magnesium chloride (MgCl2), iodine, pyrrole, dichloromethane, p-chloranil, benzaldehyde, ascorbic acid, monobasic and dibasic phosphate buffer.
A method for the preparation of mechanically triggered on-demand easy-to-fabricate donor-acceptor nanocomposite photocatalyst comprising the steps of
cutting raw aloe vera into small pieces and washed with DI water several times after the green skin was carefully removed;
grounding the internal part of the aloe vera leaf in the mixer along with ethanol;
stirring the prepared mixture at 550 rpm along with 150°C continues until the mixed solution becomes dry;
transferring a specific amount of aloe vera gel to a crucible and heated in a muffle furnace at 500°C for four hours after drying;
removing the crucible from the muffle furnace and allowed to cool at room temperature after the heating process;
obtaining product aloe vera drive graphene;
adding benzaldehyde (1 mmol), iodine (0.21 equiv), and pyrrole (1 mmol) in turn to 10 mL of dichloromethane (DCM);
after adding the para-chloranil (0.75 equivalent), a second activation (300 W, 35 0C) was carried out;
using DCM/petroleum ether (1:1) as the eluent, the mixture was purified by flash chromatography;
obtaining a purple solid (59 mg, 36%) (tetra-phenyl-porphyrin);
allowing the solution containing 0.25 g of tetra-phenyl-porphyrin and 0.28 g of MgCl2 in 10 mL of DMF to reflux for about 1 hour;
allowing the solution to cool after refluxing;
adding 200 mL of distilled water to the cooled solution and it was maintained in an ice bath for 15 minutes;
following filtering, the solid precipitate of Mg-TPP was obtained; and
rinsing the product with distilled water and left to dry;
dispersing Graphene powder in chloroform (CHCl3) and then sonicated at 45 kHz for an hour;
adding PLA pellets along with CHCl3 and Mg-TPP to the mixture;
stirring the mixture at 170°C for two hours and 30 minutes;
pouring the resin blend into a mould sized 100 × 100 × 3 mm and dried for 24 hours under solar light;
cutting the fabricated composite for mechanical strength testing in accordance with ASTM standards.
The testing mechanical strength in accordance with ASTM standards, followed by tensile, flexural, impact, and hardness tests, as well as 1,4-NADH regeneration.
For comparative mechanical strength testing, four types of specimens were developed: P neat, 0.5, 1, and 1.5% PGT nanocomposite flexible artificial leaf photocatalyst.
The production of formic acid from CO2 was also carried out in a quartz reactor at ambient temperature in an inert environment utilising a solar light source.
Using a solar light source and an inert atmosphere, photochemical regeneration of 1,4-NADH was carried out at ambient temperature in a quartz reactor.
EXAMPLE 1
EXPERIMENTAL SECTION
Materials. Nicotinamide adenine dinucleotide (ß-NAD+), formate dehydrogenase enzyme, 2,2'-bipyridine, (pentethylcyclopentadienyl) rhodium (III) chloride dimer, N, N-dimethylformamide, polylactic acid, chloroform, cyclohexane, silica gel, and magnesium chloride (MgCl2), iodine, pyrrole, dichloromethane, p-chloranil, benzaldehyde, ascorbic acid, monobasic and dibasic phosphate buffer, were purchased from the TCI.
Lighting the Path to Renewable Energy: Harnessing Photocatalysis for 1,4-NADH Production. Using a solar light source and an inert atmosphere, photochemical regeneration of 1,4-NADH was carried out at ambient temperature in a quartz reactor. Here is how the photocatalytic regeneration of 1,4-NADH was done. A quartz reactor was used to carry out the reaction. In 3.1 mL of sodium phosphate buffer (100 mM, pH 7.0), the following components of the reaction were present: ß-NAD+ (1.24 mmol), rhodium complex Rh (0.62 mmol) synthesised as per reported paper[56], AsA (1.24 mmol), and film photocatalyst (1×1cm2). The UV-vis spectrophotometer was used to track the regeneration of 1,4-NADH.
Transforming CO2 into Formic Acid Fuel. The production of formic acid from CO2 was also carried out in a quartz reactor at ambient temperature in an inert environment utilising a solar light source. The procedure used film photocatalyst (1×1cm2), ß-NAD+ (1.24 mmol), rhodium complex Rh (0.62 mmol), and formate dehydrogenase (3 units) in 3.1 mL of sodium phosphate buffer (100 mM, pH 7.0) with AsA (1.24 mmol) in the presence of CO2. After bubbling CO2 for 1 hour without light (light off), the reactor was exposed to visible light (light on). The quantity of formic acid was determined using UV-visible spectroscopy and HPLC.
Synthesis of Aloe-vera derived graphene (G). As shown in fig 1, raw aloe vera was cut into small pieces and washed with DI water several times after the green skin was carefully removed. Further internal part of the aloe vera leaf was grounded in the mixer along with ethanol. After that prepared mixture was stirred at 550 rpm along with 150°C. The stirring process continues until the mixed solution becomes dry. After drying, a specific amount of aloe vera gel is transferred to a crucible and heated in a muffle furnace at 500°C for four hours. After the heating process, the crucible is removed from the muffle furnace and allowed to cool at room temperature. Eventually, yielded aloe vera drive graphene.
Synthesis of light harvesting Magnesium tetraphenylporphyrin (T). The reported procedure [90] was followed in the synthesis of the tetra-phenyl-porphyrin Fig. 2. Adding benzaldehyde (1 mmol), iodine (0.21 equiv), and pyrrole (1 mmol) in turn to 10 mL of dichloromethane (DCM) was done without taking any special care. Following the initial activation (300 W, 35 0C), TLC revealed that benzaldehyde had completely converted. After adding the para-chloranil (0.75 equivalent), a second activation (300 W, 35 0C) was carried out. Using DCM/petroleum ether (1:1) as the eluent, the mixture was purified by flash chromatography. A purple solid (59 mg, 36%) (tetra-phenyl-porphyrin) was obtained. Finally, the solution containing 0.25 g of tetra-phenyl-porphyrin and 0.28 g of MgCl2 in 10 mL of DMF was allowed to reflux for about 1 hour. After refluxing, the solution was allowed to cool. After that, 200 mL of distilled water was added to the cooled solution and it was maintained in an ice bath for 15 minutes. Following filtering, the final solid precipitate of Mg-TPP was obtained. The product was rinsed with distilled water and left to dry [90,91].
Development of Aloe vera-derived PGT nanocomposite flexible artificial leaf photocatalyst.
Fig 3 depicts the fabrication of a PGT nanocomposite flexible artificial leaf photocatalyst. During the fabrication processes to prevent agglomeration, G powder was first dispersed in chloroform (CHCl3) and then sonicated at 45 kHz for an hour. P pellets were then added to the mixture, along with CHCl3, and the synthesized G powder was stirred at 170°C for two hours and 30 minutes. The resin blend was poured into a mold sized 100 × 100 × 3 mm and dried for 24 hours under solar light. The fabricated composite was cut for mechanical strength testing in accordance with ASTM standards, followed by tensile, flexural, impact, and hardness tests, as well as 1,4-NADH regeneration. For comparative mechanical strength testing, four types of specimens were developed: P neat, 0.5, 1, and 1.5% PGT nanocomposite flexible artificial leaf photocatalyst.
EXAMPLE 2
Result and discussion
Powering Tomorrow: The Sun-Fueled Symphony of Photocatalysts and Biocatalysts for Formic Acid Generation
The PGT nanocomposite photocatalyst when integrated with a biological enzyme i.e., formate dehydrogenase exhibited highly selective production of formic acid from CO2. The 'T' component (magnesium tetra-phenyl-porphyrin) of the PGT nanocomposite photocatalyst serves as the light-harvesting complex (electron donor) and collects incoming photons from the visible light. On absorbance of photon energy, electronic transition from HOMO to LUMO occurs and the electrons are conducted across the 'G' component (aloe-vera-derived graphene) which acts as a multi-electron acceptor. There exist an electron transport chain in the photocatalytic system which is shown in figure 3d. Concurrently, with the help of Rh-complex [56] the electrons are mediated to the NAD+ which is then reduced to 1,4 NADH. In this process the reduced Rh complex eliminates a proton from the aqueous solution and transfers a hydride ion to NAD+ and converts it into 1,4 NADH, thereby completing a photocatalytic cycle. This 1,4 NADH cofactor is then consumed by formate dehydrogenase enzyme that further enables in the conversion of CO2 to HCOOH. The bio-photocatalytic cycle is well-coupled, triggering formic acid production.
Characterization of PGT nanocomposite flexible film as an artificial photocatalyst
The T, G, and 1%PGT UV absorption investigations were carried out in solid media. Strong band absorption at 400-600 nm was observed in 1%PGT (Figure 4a), which is in line with that of T. There have already been reports of this kind of observation [57,58]. Around 420 nm, the absorption coefficients of T and G are low, whereas the absorption of 1%PGT is higher within the same range.
Additionally, the increment of absorbance of 1%PGT in the range of 400 nm to 700 nm shows the enhancement of p-p interactions, high molar extinction coefficient, and efficacious migration of energy and photo-induced electrons among T, G, and P. Figure 1b exhibits the band gap of T and 1%PGT as 2.68 and 2.8 eV, respectively, which is calculated by Kubelka-Munk (K-M) method. This shows that the electronic characteristics of G and P have relaxed in the ground-state attachment of the T core, demonstrating the high effectiveness of 1%PGT as a photocatalytic energy for regeneration of 1,4-NADH, which is crucial in the enzymatic formation of solar fuels/chemicals from CO2 under solar light.
Fourier transform infrared (FT-IR) spectroscopy was investigated to examine the existence of interfacial bonds formed between the moieties in the composite [59]. The spectra shown in Figure 5a explain the mode of different vibrational frequencies that arise due to molecular interaction. The interested region for G' is 1720 cm-1-1020 cm-1, i.e. shows characteristic peaks for C=O, C-O, and O-H vibrational stretching at 1721 cm-1, 1017 cm-1, and 1418 cm-1 respectively whereas for neat P it is at 1187cm-1 for C-O stretching [60,61]. However, the incorporation of 'P' into G' shows different vibrational and stretching peaks at 546 cm-1, 590 cm-1, 740 cm-1, and 843 cm-1 in given spectra in Figure 2a which confirms incorporation. Also, two distinct infrared peaks were observed for C-C-C bending and CH2 vibration at 1070 cm-1 and 1402 cm-1 respectively. Nonetheless, if 'T' was incorporated in 'P' instead of G' then due to the presence of a pyrrole unit, two characteristic peaks were observed for C=C and C=N vibrational modes with strong absorption band at 1081 cm-1 and 1746 cm-1 respectively. The N-H stretching vibration and bending vibration were found to be at 2985 cm-1 and 958 cm-1 respectively which disappear due to the insertion of metal like magnesium (in the case of metalloporphyrin). Moreover, all of the vibrations and stretching peaks in the comparative profiles of 0.5% PGT, 1% PGT, and 1.5% PGT correspond to a decrease in transmittance intensity. This FTIR assessment confirms that G and T react with P and produces a new PGT nanocomposite flexible artificial leaf photocatalyst which helps improve the cohesion strength and adhesion strength with P. The above study also revealed that reinforcement of G and T in the P composites improves the tensile, flexural, impact strength, and hardness of the PGT nanocomposite flexible artificial leaf photocatalyst [62].
X-ray diffraction (XRD) analysis was performed on a developed sample, including G, P neat, P@G, P@T, and a 1% PGT nanocomposite flexible film as an artificial leaf photocatalyst. In Figure 5b, G showed multiple peaks, which is consistent with previous research, which shows that graphene typically has a weak broad peak at 26.12° [63-65]. The P@G sample exhibited two broader peaks at 16.98° and 19.32°, indicating the presence of P and G. The P neat and P@T material showed broad peaks as per the reported method [66]. The 1% PGT nanocomposite flexible artificial leaf photocatalyst, noted for its exceptionally high strength, showed additional peaks at 16.84°, 20.01°, and 22.54°, confirming the presence of G and T in this newly developed nanocomposite photocatalyst. As a result, the 1% PGT nanocomposite flexible artificial leaf photocatalyst outperformed the other samples, indicating its potential for a variety of applications.
TGA plays an essential role in better understanding the thermal properties of the composite [67]. Therefore, it can help to investigate the thermal degradation and thermal stability of P neat, P@G, and 1%PGT nanocomposite flexible film as an artificial leaf photocatalyst. The weight loss vs temperature curve that was acquired from TGA was used to analyze the thermal stability of the 1%PGT filament. Figure 2c illustrates how weight loss in the composite sample starts to happen at temperatures over 1800C, which is when thermal degradation starts. About 1% and 4% of the P and PGT composite's weight was lost between 1100C and 3000C, respectively. A mass loss of almost 95% was seen for the composite sample (compared to a 100% loss for pure P) when the samples were heated from room temperature to 750 0C in an N2 environment, confirming G and T materials. From the TGA plot of the residue, it is obvious that G and T are the principal filler components included in the nanocomposite photocatalyst. For the P and PGT nanocomposite flexible film as artificial leaf photocatalysts, weight loss was approximately 100% and 95% between 4000C and 3500C, respectively. This suggests that the PGT nanocomposite homogeneity caused the composite to degrade gradually and under control [68,69] .
Electroanalytical approach for determination of band gap and formation of PGT nanocomposite flexible film as an artificial leaf photocatalyst
The redox potential of T and 1% PGT nanocomposite flexible artificial leaf photocatalyst is examined by cyclic voltammetry (CV) with glassy carbon electrode (working electrode), calomel electrode (reference electrode), and platinum electrode (Counter electrode) in 0.1M tetrabutylammonium hexafluorophosphate (TBAHFP) electrolyte[70]. The cyclic voltammetry experiment is performed to demonstrate the oxidation and reduction potential of T and 1% PGT nanocomposite flexible artificial leaf photocatalyst. In Figure 6a, the oxidation potential was observed at an anodic peak of T (+1.29V) and a cathodic peak (-1.24V). The oxidation and reduction potential of 1% PGT was observed at an anodic peak (+1.22V) and cathodic peak (-1.20V). The 1% PGT showed an anodic shift of oxidation potential, which designates the effective migration of charge carriers from the 1% PGT to Rh complex. The photoelectrical property of 1% PGT originates from the excitation of electrons from HOMO (5.72eV) to the LUMO (3.30eV), followed by the transfer of the excited electrons into the Rh complex (Figure 6d) [71] . The vicinity and potential gradient between the light-harvesting and Rh center enable effective electron transfer from the 1% PGT photocatalyst to the Rh. Further, on comparing the band gaps of T (2.59 eV) and 1% PGT (2.42 eV), the 1% PGT photocatalyst has a narrower band gap, allowing electrons from the excited 1% PGT to easily move to the Rh complex. This finding is consistent with band gap predictions using the Tauc Method, demonstrating the effective electron transport mechanism from the 1%PGT to the Rh complex in this photocatalytic system. Tafel plot elucidates the photocatalytic activity of T and 1% PGT nanocomposite flexible artificial leaf photocatalyst.
The Tafel slope on a Tafel plot is inversely associated with the photocatalytic activity of the material. A high slope denotes a rapid reaction rate, whereas a shallow slope indicates a slower rate. In photocatalysis, a steep slope would indicate effective electron-hole pair formation and separation, which leads to high photocatalytic activity. Figure 3b, shows the Tafel plot of T and 1% PGT photocatalyst. The Tafel slope of 1% PGT is lower than the T indicating the better photocatalytic activity of our newly designed 1% PGT.
Electrochemical impedance spectroscopy (EIS) is an effective method for studying charge carrier dynamics in a three-electrode system. Our study included EIS analysis at frequencies ranging from 100kHz to 0.1Hz. The AC amplitude was tuned to 5mV to study charge transfer and transport in the system. In order to ensure a uniform testing environment, we maintain a working/reference electrode separation of 1 cm during electrochemical studies [70]. Figure 6c shows Nyquist plots of P@G and 1%PGT photocatalyst. Nyquist plot consists of a plot of the imaginary part (-Z')versus the real part (Z') of the impedance. The plot frequently takes the shape of a semicircle, and the arc radius of this semicircle is proportional to the charge transfer resistance of the system. Smaller the semicircle arc radius, lower the charge transfer resistance which leads to high charge transfer. Figure 3c shows small arc radius for 1%PGT as compared to P@G. Which clearly evidence that the newly synthesized 1%PGT photocatalyst has better charge carrier dynamics due to lower charge transfer resistance than the P@G.
A plausible mechanism of transference of photo-generated charge within the 1%PGT photocatalyt has been demonstrated in figure 6d. Upon irradiation under visible light, the electrons from the valence band(-5.72eV) of the 1%PGT photocatalyst jumps to the conduction band(-3.30eV) generating electron-hole pairs. Ascorbic acid (AsA) is acting as a sacrificial agent which quenches the VB of the 1%PGT photocatalyst. The electrons from the conduction band(CB) is then transferred to the NAD+(-4.2eV) with the help of Rh complex electron mediator(-3.96eV).This movement of electrons from the photocatalyst to the NAD+ leads to the formation of 1,4 NADH cofactor [72].
Unveiling the Morphological Marvels of Photocatalysts
Field emission scanning electron microscopy (FE-SEM) is a morphological research tool for capturing the microstructure image of developed materials. Figure 6 (a-d) depicts the FE-SEM images of P neat, P@G, P@T, and 1% PGT nanocomposite flexible artificial leaf photocatalyst. Figure 4a depicts the structure of P neat [73], while the yellow circle in (b) indicates the presence of G content in the P matrix [74]. Figure 4c depicts the presence of T content in the P matrix in the form of a crumbling structure, while Figure 6d shows the presence of G and T particles in the matrix. It can be seen that G and T particles have joints and protuberances attached to the surface of P, which helps in mechanical interlocking and reduces fatigue crack propagation in the composite. Overall, the PGT nanocomposite photocatalyst exhibits good dispersion.
Mechanical studies of designed PGT nanocomposite flexible film as an artificial leaf photocatalyst.
[A] Study of tensile strength of designed materials: The mechanical properties of the newly developed nanocomposite photocatalyst were determined using tensile tests. These tests entail the application of force to a material until it reaches its breaking point, thereby enabling the determination of its tensile strength and elongation properties (Figure 8). The specimens, with dimensions of 63.5 mm × 19 mm × 3 mm, were tested using a universal testing machine (UTMModel-EZ50) with a maximum capacity of 50KN at room temperature, following the guidelines set by ASTM D638 standards.
Tensile tests were conducted on nanocomposite photocatalyst specimens with varying concentrations of PGT (0%, 0.5%, 1%, and 1.5%), as depicted in Figure 6. For each variation, three specimens are tested, and the average outcomes have been recorded which is presented in Table 1. Figure 9a-b depicts the specimens before and after testing, demonstrating the successful tensile tests. Moreover, Figure 10 illustrates the stress-strain characteristics of the tensile samples, offering insights into their mechanical efficacy. The experimental data demonstrates that incorporating G reinforcement consistently enhances tensile strength by up to 1%. This enhancement is due to G and T developing strong cross-links with the matrix, delaying crack initiation and propagation [75,76]. However, after a certain amount of fraction, the load-carrying capacity decreases due to G agglomeration. This phenomenon emphasizes the critical importance of optimizing the G content in the nanocomposite to achieve optimal mechanical properties. Therefore, the optimum value for maximum tensile strength is obtained for 1 % PGT nanocomposite flexible artificial leaf photocatalyst, and the value of maximum load capacity is 589.49 N. This highly tensile nanocomposite photocatalyst can be applied to scaffolding applications in human tissue engineering.
Table. 1 Mechanical properties of developed nanocomposite flexible artificial leaf photocatalyst.
S. No. Developed specimen code Composition of developed specimen Tensile properties
Tensile strength (MPa) Tensile load (N) Strain (%)
1. 0 % PGT P, 0% G, T 16.184 307.30 4.761
2. 0.5 % PGT P, 0.5% G, T 17.125 432.58 6.771
3. 1.0 % PGT P, 1% G, T 25.322 589.49 11.755
4. 1.5 % PGT P, 1.5% G, T 18.292 349.56 10.878
[B] Flexural Test: A flexural test, also known as a bending test, is a material testing method used to determine the behaviour of a material when subjected to bending loads [77]. This test is particularly important in assessing the strength and stiffness of materials, especially those used in structural applications like construction materials, metals, plastics, and composites [78]. As shown in Figure 8, the flexural test has been performed as per ASTM standard ASTM D790 (125×12.7×3.2) on the developed newly PGT nanocomposite flexible artificial leaf photocatalyst [79]. For each composition, three specimens were examined under a three-point bending mode. Figure 11 shows the stress-strain behaviour of 0 %, 0.5 %, 1%, and 1.5% PGT nanocomposite flexible artificial leaf photocatalyst. The calculated flexural strength and load are detailed in Table 2, which shows that reinforcement of G and T in P composites improves flexural strength. Notably, the highest flexural strength was found at 1% PGT nanocomposites flexible film as an artificial leaf photocatalyst. However, the use of nanomaterials promotes agglomeration, which serves as a site for crack propagation. [80-82]
Table. 2 Flexural test results of developed nanocomposite photocatalyst
S.No Developed specimen code Composition of developed specimen Flexural Properties
Maximum bending Stress (MPa) Maximum bending Strain (%) Tensile load (N)
1 0 % PGT P, 0% G, T 30.019 6.2054 85.835
2 0.5 % PGT P, 0.5% G, T 31.196 15.787 106.91
3 1 % PGT P, 1% G, T 32.244 17.18 163.01
4 1.5 % PGT P, 1.5% G, T 31.577 17.513 138.23

[C] Impact Testing: The impact test evaluates the materials' energy absorption and impact resistance during the collision phase of the nanocomposite specimen [83]. The Izod impact test used an interchangeable pendulum-type hammer Izod impact tester machine with a pendulum speed of 346 m/s. A pivoting arm is raised to a specific height and then released. As shown in Figure 12a, the striker swings down, hitting a notched sample and breaking the specimen. The Izod impact test is performed according to the ASTM D256 standard with a 64 × 12.7 × 3 mm specimen size [84].
The role of G and T content on the impact behaviour of developed nanocomposite photocatalysts is shown in Figure 13b. The results demonstrated that the impact strength of P neat composite could absorb energy up to 1.4115 J. In contrast, reinforcement of G and T in P neat composite increases the energy absorption capacity. It can be seen that the maximum impact strength of 2.461 J is obtained by 1 % PGT nanocomposite flexible artificial leaf photocatalyst. The impact strength of 0.5 % and 1.5 % of the PGT nanocomposite photocatalyst are shown at 1.854 J and 1.937 J. The developed nanocomposite with 0.5 and 1.5% PGT reinforcement shows a slightly better enhancement than the P neat composite. The overall 23 %, 42.65 %, and 27.129 % improvement was observed in impact strength at 0.5, 1, and 1.5 PGT nanocomposite flexible artificial leaf photocatalyst compared to the P neat composite. A decreasing trend in impact strength was obtained after a fraction of the reinforcement of G in the P composite. It may be due to the agglomeration effect in matrix resin, which propagates the pre-cracking in the polymer [85]. During this experiment, the matrix initially absorbs the impact load. The strength of the matrix was enhanced through modifications in its G and T content. Incorporating additional G up to 1.5 % fractions did not change the results noticeably, indicating that the matrix plays a prominent role in the composite capacity to absorb impact energy [86]. In this impact strength study, the G and T play a vital role in enhancing the impact strength and increasing the energy-absorbing capacity.
[D] Shore D Hardness: A Shore D durometer tester determined the newly developed nanocomposite photocatalyst's surface hardness. Shore D hardness measures a composite's resistance to needle penetration at a given spring force [69,87]. The hardness is graded on a scale of 0 to 100 and classified as A or D. In this test, a higher numerical value indicates greater hardness, with A denoting flexibility and D indicating rigidity. The shore D hardness test involves testing three specimens of each developed nanocomposite (P neat, 0.5%, 1%, and 1.5 PGT nanocomposite flexible artificial leaf photocatalyst, Fig 14) and calculating the average value. According to the literature review, hardness test results influence mechanical properties such as shear strength, flexural strength, tensile strength, modulus of elasticity, and impact strength. Figure 11 depicts the effect of reinforcement P and T content on the hardness of the nanocomposite photocatalyst. The hardness value of PGT nanocomposite flexible artificial leaf photocatalyst is observed to increase from 54.52 J (P neat) to 60.4, 65.08, and 63.96 with reinforcement of 0.5, 1, and 1.5 Wt.% PGT. The optimal results obtained with 1 Wt.% PGT nanocomposite photocatalysts are due to well-developed interfacial interaction and enriched G dispersion. Compared to the P neat composite, hardness improved by 10.78%, 19.36%, and 17.31% at 0.5, 1, and 1.5 Wt.% PGT nanocomposite flexible artificial leaf photocatalyst, respectively. The results of this investigation indicate that the reinforcement of G and T in P improves mechanical properties.
Green Alchemy: Harnessing Light for 1,4-NADH Regeneration and Formic Acid Synthesis from CO2. A series of photocatalysis experiments were performed to study the photocatalytic activities of P, T, 0.5%PGT, 1% PGT, and 1.5% PGT for the visible light-driven photoregeneration of 1,4-NADH. The concentration of photogenerated 1,4-NADH was measured by a spectrophotometer [88]. As shown in Figure 12, 1% PGT nanocomposite flexible film as an artificial leaf is significantly effective for 1,4-NADH photoregeneration constantly accumulating up to 61.09% with a time linearity. However P, T, 0.5%PGT and 1.5% PGT photocatalyst gives only 0.00%, 9.27% 30.54% and 60.91% yield of 1,4-NADH regeneration (Figure 15a). Additional experiments were performed to study the photocatalytic performance of P, T, 0.5%PGT, 1% PGT, and 1.5% PGT for the visible-light-driven artificial photosynthesis of formic acid from CO2. The amount of formic acid was detected by UV-visible spectroscopy and HPLC (Model:LC-20AP). As shown in Figure 15b, the formic acid yield increases linearly with the reaction time when 1% PGT nanocomposite flexible film as an artificial leaf is used as a photocatalyst. The efficiency of 1% PGT nanocomposite flexible film as an artificial leaf for the production of formic acid for 2 h was 146.62 µmol, while P, T, 0.5%PGT, and 1.5% PGT were 0.00, 22.24, 73.31 and 122.18 µmol, respectively. These investigations clearly reveal the superiority of the 1% PGT nanocomposite flexible film as an artificial leaf photocatalyst over the other photocatalysts due to superior mechanical properties.
The incident light is absorbed at T, where photoexcitation occurs, and the created electrons move into the P via a stable mechanical bonding such as bending strength, tensile strength, shear strength etc. The energy level alignment between the lowest unoccupied molecular orbital (LUMO) of T and the conduction band edges of P@G could be used to assess the electron-transfer efficiency.
Ascorbic acid (AsA) is used as a hole scavenger to prevent electron transport from P@G to the photoexcited T. Therefore, rather than the other way around, it is simple to move the photoexcited electrons produced in T into P@G. The possibility of electrons being transported into the rhodium complex is increased by the high surface active site and exceptionally high carrier mobility of P@G [89] thereby speeding up the chemical processes leading to the formation of 1,4-NADH. In addition, the P@G transports multiple electrons by serving as an electron reservoir. Due to strong Coulomb repulsion between localised electrons, electron addition in most molecular systems has a very high energy cost. In P@G, the delocalized nature of wave functions throughout the whole P@G surface (extending many µm) makes multielectron addition or removal conceivable.
Conclusions
Undoubtedly, the creation of an artificial leaf made of 1% PGT photocatalyst integrated donor-acceptor nanocomposite is an impressive accomplishment. This nanocomposite has remarkable capabilities as a film photocatalyst, effectively utilising solar light to propel the regeneration of 1,4-NADH and the subsequent synthesis of formic acid from CO2 via the activity of the enzyme formate dehydrogenase. The regeneration efficiency of 61.09 ± 0.59% and the production of formic acid of roughly 146.62 ± 1.6 µmol demonstrate how well this system works with renewable energy sources to carry out useful chemical reactions. In addition, the PGT nanocomposite exhibits equally remarkable mechanical characteristics, including high strain (11.755%), bending strength (32.244MPa), tensile strength (25.322MPa), tensile load (589.49 Newton), and impact energy (2.4615 Joule). Because of these qualities, the nanocomposite is a good material for a variety of uses, such as tissue implants, where biocompatibility, flexibility, and durability are essential. This PGT nanocomposite, which combines exceptional mechanical and photocatalytic capabilities, is a major breakthrough in materials research and has exciting prospects for biomedical engineering and sustainable energy conversion. Absolutely, the comparison highlights the superiority of the 1% PGT nanocomposite flexible artificial leaf as a film photocatalyst in CO2 fixation into formic acid (HCOOH) compared to its counterparts with 0.5% and 1.5% PGT. This highlights the crucial function of P-based nanocomposite artificial leaves (PGT) in a number of fields, such as scaffold-based tissue engineering, photocatalysis, and solar chemical synthesis. The 1% PGT nanocomposite's superior performance over other compositions highlights the well-balanced combination of its mechanical qualities and photocatalytic activity. This ideal balance emphasises the significance of fine-tuning material compositions for particular applications and is essential for effective CO2 conversion. All things considered, the findings highlight the versatility and promise of PGT nanocomposites in a variety of domains, from medicinal applications to sustainable energy conversion. They serve as a flexible framework for tackling urgent issues in fields like tissue engineering and carbon capture and utilisation, demonstrating the wide-ranging influence of materials innovation on solving global issues.
Non Patent References
[1] G. Wang, L. Zhang, J. Zhang, A review of electrode materials for electrochemical supercapacitors, Chem. Soc. Rev. 41 (2012) 797-828. https://doi.org/10.1039/c1cs15060j.
[2] P.A.T. Kelly, C.J. Davis, G.M. Goodwin, Differential patterns of local cerebral glucose utilization in response to 5-hydroxytryptamine, agonists, Neuroscience. 25 (1988) 907-915. https://doi.org/10.1016/0306-4522(88)90044-9.
[3] S.S. Balaji, M. Sathish, Supercritical fluid processing of nitric acid treated nitrogen doped graphene with enhanced electrochemical supercapacitance, RSC Adv. 4 (2014) 52256-52262. https://doi.org/10.1039/c4ra07820a.
[4] B. Amutha, M. Sathish, A 2 V asymmetric supercapacitor based on reduced graphene oxide-carbon nanofiber-manganese carbonate nanocomposite and reduced graphene oxide in aqueous solution, J. Solid State Electrochem. 19 (2015) 2311-2320.
[5] K. Subramani, D. Jeyakumar, M. Sathish, Manganese hexacyanoferrate derived Mn 3 O 4 nanocubes--reduced graphene oxide nanocomposites and their charge storage characteristics in supercapacitors, Phys. Chem. Chem. Phys. 16 (2014) 4952-4961.
[6] K. Subramani, N. Lakshminarasimhan, P. Kamaraj, M. Sathish, Facile and scalable route to sheets-on-sheet mesoporous Ni--Co-hydroxide/reduced graphene oxide nanocomposites and their electrochemical and magnetic properties, RSC Adv. 6 (2016) 15941-15951.
[7] E. Frackowiak, Carbon materials for supercapacitor application, Phys. Chem. Chem. Phys. 9 (2007) 1774-1785.
[8] R.J. Mo, Y. Zhao, M. Wu, H.M. Xiao, S. Kuga, Y. Huang, J.P. Li, S.Y. Fu, Activated carbon from nitrogen rich watermelon rind for high-performance supercapacitors, RSC Adv. 6 (2016) 59333--59342, (n.d.).
[9] F.-C. Wu, R.-L. Tseng, C.-C. Hu, C.-C. Wang, Physical and electrochemical characterization of activated carbons prepared from firwoods for supercapacitors, J. Power Sources. 138 (2004) 351-359.
[10] F.-C. Wu, R.-L. Tseng, C.-C. Hu, C.-C. Wang, Effects of pore structure and electrolyte on the capacitive characteristics of steam-and KOH-activated carbons for supercapacitors, J. Power Sources. 144 (2005) 302-309.
[11] P. Sharma, T.S. Bhatti, A review on electrochemical double-layer capacitors, Energy Convers. Manag. 51 (2010) 2901-2912.
[12] B.E. Conway, V. Birss, J. Wojtowicz, The role and utilization of pseudocapacitance for energy storage by supercapacitors, J. Power Sources. 66 (1997) 1-14.
[13] L.L. Zhang, X.S. Zhao, Carbon-based materials as supercapacitor electrodes, Chem. Soc. Rev. 38 (2009) 2520-2531.
[14] D. Puthusseri, V. Aravindan, S. Madhavi, S. Ogale, 3D micro-porous conducting carbon beehive by single step polymer carbonization for high performance supercapacitors: the magic of in situ porogen formation, Energy \& Environ. Sci. 7 (2014) 728-735.
[15] S. Bai, G. Tan, X. Li, Q. Zhao, Y. Meng, Y. Wang, Y. Zhang, D. Xiao, Pumpkin-derived porous carbon for supercapacitors with high performance, Chem. Asian J. 11 (2016) 1828-1836.
[16] Y.-Y. Wang, B.-H. Hou, H.-Y. Lü, C.-L. Lü, X.-L. Wu, Hierarchically Porous N-Doped Carbon Nanosheets Derived From Grapefruit Peels for High-Performance Supercapacitors, ChemistrySelect. 1 (2016) 1441-1447.
[17] T.E. Rufford, D. Hulicova-Jurcakova, K. Khosla, Z. Zhu, G.Q. Lu, Microstructure and electrochemical double-layer capacitance of carbon electrodes prepared by zinc chloride activation of sugar cane bagasse, J. Power Sources. 195 (2010) 912-918.
[18] M. Wahid, D. Puthusseri, D. Phase, S. Ogale, Enhanced capacitance retention in a supercapacitor made of carbon from sugarcane bagasse by hydrothermal pretreatment, Energy \& Fuels. 28 (2014) 4233-4240.
[19] T.E. Rufford, D. Hulicova-Jurcakova, Z. Zhu, G.Q. Lu, Nanoporous carbon electrode from waste coffee beans for high performance supercapacitors, Electrochem. Commun. 10 (2008) 1594-1597.
[20] M.S. Balathanigaimani, W.-G. Shim, M.-J. Lee, C. Kim, J.-W. Lee, H. Moon, Highly porous electrodes from novel corn grains-based activated carbons for electrical double layer capacitors, Electrochem. Commun. 10 (2008) 868-871.
[21] X. Wu, W. Xing, J. Florek, J. Zhou, G. Wang, S. Zhuo, Q. Xue, Z. Yan, F. Kleitz, On the origin of the high capacitance of carbon derived from seaweed with an apparently low surface area, J. Mater. Chem. A. 2 (2014) 18998-19004.
[22] J. Wang, S. Kaskel, KOH activation of carbon-based materials for energy storage, J. Mater. Chem. 22 (2012) 23710-23725.
[23] Q. Cao, K.-C. Xie, Y.-K. Lv, W.-R. Bao, Process effects on activated carbon with large specific surface area from corn cob, Bioresour. Technol. 97 (2006) 110-115.
[24] J. Sahira, A. Mandira, P.B. Prasad, P.R. Ram, Effects of activating agents on the activated carbons prepared from lapsi seed stone, Res. J. Chem. Sci. 2231 (2013) 606X.
[25] M. Sathish, S. Mitani, T. Tomai, I. Honma, Supercritical fluid assisted synthesis of N-doped graphene nanosheets and their capacitance behavior in ionic liquid and aqueous electrolytes, J. Mater. Chem. A. 2 (2014) 4731-4738.
[26] M. Beidaghi, Y. Gogotsi, Capacitive energy storage in micro-scale devices: recent advances in design and fabrication of micro-supercapacitors, Energy \& Environ. Sci. 7 (2014) 867-884.
[27] R. Currie, S. Mottaghi-Tabar, Y. Zhuang, D.S.A. Simakov, Design of an air-cooled sabatier reactor for thermocatalytic hydrogenation of CO2: experimental proof-of-concept and model-based feasibility analysis, Ind. \& Eng. Chem. Res. 58 (2019) 12964-12980.
[28] D. Ferrah, A.R. Haines, R.P. Galhenage, J.P. Bruce, A.D. Babore, A. Hunt, I. Waluyo, J.C. Hemminger, Wet chemical growth and thermocatalytic activity of Cu-based nanoparticles supported on TiO2 nanoparticles/HOPG: in situ ambient pressure XPS study of the CO2 hydrogenation reaction, Acs Catal. 9 (2019) 6783-6802.
[29] G. Takalkar, R.R. Bhosale, F. AlMomani, M. Khraisheh, Thermocatalytic splitting of CO2 using sol-gel synthesized Co-ferrite redox materials, Fuel. 257 (2019) 115965.
[30] Y. Song, X. Zhang, K. Xie, G. Wang, X. Bao, High-temperature CO2 electrolysis in solid oxide electrolysis cells: developments, challenges, and prospects, Adv. Mater. 31 (2019) 1902033.
[31] B.M. Tackett, E. Gomez, J.G. Chen, Net reduction of CO2 via its thermocatalytic and electrocatalytic transformation reactions in standard and hybrid processes, Nat. Catal. 2 (2019) 381-386.
[32] A. Lisovskaya, D.M. Bartels, Reduction of CO2 by hydrated electrons in high temperature water, Radiat. Phys. Chem. 158 (2019) 61-63.
[33] D.C. Grills, J.A. Farrington, B.H. Layne, S. V Lymar, B.A. Mello, J.M. Preses, J.F. Wishart, Mechanism of the formation of a Mn-based CO2 reduction catalyst revealed by pulse radiolysis with time-resolved infrared detection, J. Am. Chem. Soc. 136 (2014) 5563-5566.
[34] J. Grodkowski, P. Neta, Copper-catalyzed radiolytic reduction of CO2 to CO in aqueous solutions, J. Phys. Chem. B. 105 (2001) 4967-4972.
[35] X. Liu, F. Kang, C. Hu, L. Wang, Z. Xu, D. Zheng, W. Gong, Y. Lu, Y. Ma, J. Wang, A genetically encoded photosensitizer protein facilitates the rational design of a miniature photocatalytic CO2-reducing enzyme, Nat. Chem. 10 (2018) 1201-1206.
[36] A. Efrati, C.-H. Lu, D. Michaeli, R. Nechushtai, S. Alsaoub, W. Schuhmann, I. Willner, Assembly of photo-bioelectrochemical cells using photosystem I-functionalized electrodes, Nat. Energy. 1 (2016) 1-8.
[37] P. Wang, S. Wang, H. Wang, Z. Wu, L. Wang, Recent progress on photo-electrocatalytic reduction of carbon dioxide, Part. \& Part. Syst. Charact. 35 (2018) 1700371.
[38] M.J. Kang, C.W. Kim, A.U. Pawar, H.G. Cha, S. Ji, W.-B. Cai, Y.S. Kang, Selective alcohol on dark cathodes by photoelectrochemical CO2 valorization and their in situ characterization, ACS Energy Lett. 4 (2019) 1549-1555.
[39] U. Kang, S.H. Yoon, D.S. Han, H. Park, Synthesis of aliphatic acids from CO2 and water at efficiencies close to the photosynthesis limit using mixed copper and iron oxide films, ACS Energy Lett. 4 (2019) 2075-2080.
[40] C.-H. Huang, C.-S. Tan, others, A review: CO2 utilization, Aerosol Air Qual. Res. 14 (2014) 480-499.
[41] A. Meng, L. Zhang, B. Cheng, J. Yu, TiO2--MnO x--Pt hybrid multiheterojunction film photocatalyst with enhanced photocatalytic CO2-reduction activity, ACS Appl. Mater. \& Interfaces. 11 (2018) 5581-5589.
[42] C. Kim, K.M. Cho, A. Al-Saggaf, I. Gereige, H.-T. Jung, Z-scheme photocatalytic CO2 conversion on three-dimensional BiVO4/carbon-coated Cu2O nanowire arrays under visible light, ACS Catal. 8 (2018) 4170-4177.
[43] S. Wang, M. Xu, T. Peng, C. Zhang, T. Li, I. Hussain, J. Wang, B. Tan, Porous hypercrosslinked polymer-TiO2-graphene composite photocatalysts for visible-light-driven CO2 conversion, Nat. Commun. 10 (2019) 676.
[44] Z. Sun, T. Ma, H. Tao, Q. Fan, B. Han, Fundamentals and challenges of electrochemical CO2 reduction using two-dimensional materials, Chem. 3 (2017) 560-587.
[45] N. Zhang, R. Long, C. Gao, Y. Xiong, Recent progress on advanced design for photoelectrochemical reduction of CO2 to fuels, Sci. China Mater. 61 (2018) 771-805.
[46] I. Ganesh, others, Conversion of carbon dioxide to methanol using solar energy-a brief review, Mater. Sci. Appl. 2 (2011) 1407.
[47] L. Wang, W. Chen, D. Zhang, Y. Du, R. Amal, S. Qiao, J. Wu, Z. Yin, Surface strategies for catalytic CO 2 reduction: from two-dimensional materials to nanoclusters to single atoms, Chem. Soc. Rev. 48 (2019) 5310-5349.
[48] S. Xie, Q. Zhang, G. Liu, Y. Wang, Photocatalytic and photoelectrocatalytic reduction of CO 2 using heterogeneous catalysts with controlled nanostructures, Chem. Commun. 52 (2016) 35-59.
[49] J. Sarkar, S. Bhattacharyya, Application of graphene and graphene-based materials in clean energy-related devices Minghui, Arch. Thermodyn. 33 (2012) 23-40. https://doi.org/10.1002/er.
[50] S. Singh, R.K. Yadav, T.W. Kim, C. Singh, P. Singh, A. P. Singh, A.K. Singh, A.K. Singh, J.O. Baeg, S.K. Gupta, Design of a graphitic carbon nitride catalytic-biocatalytic system for solar light-based CO2 production, React. Chem. Eng. 7 (2022) 1566-1572. https://doi.org/10.1039/d2re00079b.
[51] H.L. Nguyen, A. Alzamly, Covalent Organic Frameworks as Emerging Platforms for CO2Photoreduction, ACS Catal. 11 (2021) 9809-9824. https://doi.org/10.1021/acscatal.1c02459.
[52] M. Khalil, J. Gunlazuardi, T.A. Ivandini, A. Umar, Photocatalytic conversion of CO2 using earth-abundant catalysts: A review on mechanism and catalytic performance, Renew. Sustain. Energy Rev. 113 (2019) 109246.
[53] R.J. Detz, B. van der Zwaan, Transitioning towards negative CO2 emissions, Energy Policy. 133 (2019) 110938.
[54] S. Kar, A. Goeppert, G.K.S. Prakash, Combined CO2 capture and hydrogenation to methanol: amine immobilization enables easy recycling of active elements, ChemSusChem. 12 (2019) 3172-3177.
[55] R.K. Yadav, J.O. Lee, A. Kumar, N.J. Park, D. Yadav, J.Y. Kim, J.O. Baeg, Highly Improved Solar Energy Harvesting for Fuel Production from CO2 by a Newly Designed Graphene Film Photocatalyst, Sci. Rep. 8 (2018) 1-10. https://doi.org/10.1038/s41598-018-35135-7.
[56] S. Chaubey, R.K. Yadav, S.K. Tripathi, B.C. Yadav, S.N. Singh, T.W. Kim, Covalent Triazine Framework as an Efficient Photocatalyst for Regeneration of NAD(P)H and Selective Oxidation of Organic Sulfide, Photochem. Photobiol. 98 (2022) 150-159. https://doi.org/10.1111/php.13504.
[57] R.K. Yadav, G.H. Oh, N.J. Park, A. Kumar, K.J. Kong, J.O. Baeg, Highly selective solar-driven methanol from CO2 by a photocatalyst/biocatalyst integrated system, J. Am. Chem. Soc. 136 (2014) 16728-16731. https://doi.org/10.1021/ja509650r.
[58] R.K. Yadav, J.O. Baeg, G.H. Oh, N.J. Park, K.J. Kong, J. Kim, D.W. Hwang, S.K. Biswas, A photocatalyst-enzyme coupled artificial photosynthesis system for solar energy in production of formic acid from CO 2, J. Am. Chem. Soc. 134 (2012) 11455-11461. https://doi.org/10.1021/ja3009902.
[59] S. Kumar, R.K. Yadav, K. Ram, A. Aguiar, J. Koh, A.J.F.N. Sobral, Graphene oxide modified cobalt metallated porphyrin photocatalyst for conversion of formic acid from carbon dioxide, J. CO2 Util. 27 (2018) 107-114. https://doi.org/10.1016/j.jcou.2018.07.008.
[60] A. Sotoudeh, G. Darbemamieh, V. Goodarzi, S. Shojaei, A. Asefnejad, Tissue engineering needs new biomaterials: Poly(xylitol-dodecanedioic acid)-co-polylactic acid (PXDDA-co-PLA) and its nanocomposites, Eur. Polym. J. 152 (2021) 110469. https://doi.org/10.1016/j.eurpolymj.2021.110469.
[61] M.W. Nugraha, M.D.H. Wirzal, F. Ali, L. Roza, N.S. Sambudi, Electrospun polylactic acid/ tungsten oxide/ amino-functionalized carbon quantum dots (PLA/WO3/N-CQDs) fibers for oil/water separation and photocatalytic decolorization, J. Environ. Chem. Eng. 9 (2021) 106033. https://doi.org/10.1016/j.jece.2021.106033.
[62] M. Genedy, R. Chennareddy, E.M. Soliman, U.F. Kandil, M.M.R. Taha, Improving shear strength of bolted joints in pultruded glass fiber reinforced polymer composites using carbon nanotubes, J. Reinf. Plast. Compos. 36 (2017) 958-971. https://doi.org/10.1177/0731684417697827.
[63] R.K. Shukla, R.K. Yadav, V.L. Gole, C.Y. Na, G.H. Jeong, S. Singh, J.O. Baeg, M.Y. Choi, N.K. Gupta, T.W. Kim, Aloe vera-derived graphene-coupled phenosafranin photocatalyst for generation and regeneration of ammonia and NADH by mimicking natural photosynthetic route, Photochem. Photobiol. (2023) 1-11. https://doi.org/10.1111/php.13831.
[64] T. Lalire, A. Taguet, J.-C. Roux, B. Otazaghine, C. Longuet, Chemical modification strategies for the control of graphene localization in PS/PMMA blend, FlatChem. 39 (2023) 100500.
[65] K. Ghosal, P. Mondal, S. Bera, S. Ghosh, Graphene family nanomaterials-opportunities and challenges in tissue engineering applications, FlatChem. 30 (2021) 100315.
[66] M. Keramati, I. Ghasemi, M. Karrabi, H. Azizi, M. Sabzi, Incorporation of surface modified graphene nanoplatelets for development of shape memory PLA nanocomposite, Fibers Polym. 17 (2016) 1062-1068. https://doi.org/10.1007/s12221-016-6329-7.
[67] M. Dehghan, I. Sbarski, Thermo-Mechanical Characterization of MWCNTs-Modified Epoxy Resin, Int. J. Chem. Nucl. Metall. Mater. Eng. 8 (2014) 119-124.
[68] H. Sun, B. Yu, J. Han, J. Kong, L. Meng, F. Zhu, Microstructure, thermal properties and rheological behavior of PLA/PCL blends for melt-blown nonwovens, Polym. 38 (2014) 477-483. https://doi.org/10.7317/pk.2014.38.4.477.
[69] H. Norazlina, Y. Kamal, Elucidating the plasticizing effect on mechanical and thermal properties of poly(lactic acid)/carbon nanotubes nanocomposites, Polym. Bull. 78 (2021) 6911-6933. https://doi.org/10.1007/s00289-020-03471-2.
[70] B. Subramanya, D.K. Bhat, Novel one-pot green synthesis of graphene in aqueous medium under microwave irradiation using a regenerative catalyst and the study of its electrochemical properties, New J. Chem. 39 (2015) 420-430. https://doi.org/10.1039/c4nj01359j.
[71] K. Sharma, R.K. Yadav, S. Singh, C. Singh, R. Shahin, K. Kumar, R.K. Verma, D.K. Dwivedi, J.O.K. Baeg, N.K. Gupta, Sun-powered CO2 transformation: TBE-Y photocatalyst's remarkable selectivity for solar-induced HCOOH, Mol. Catal. 548 (2023) 113379. https://doi.org/10.1016/j.mcat.2023.113379.
[72] S. Chaubey, P. Singh, C. Singh, S. Singh, S. Shreya, R.K. Yadav, S. Mishra, Y. Jeong, B.K. Biswas, T.W. Kim, Ultra-efficient synthesis of bamboo-shape porphyrin framework for photocatalytic CO2 reduction and consecutive CS/CN bonds formation, J. CO2 Util. 59 (2022) 101968.
[73] M. Gong, Q. Zhao, L. Dai, Y. Li, T. Jiang, Fabrication of polylactic acid/hydroxyapatite/graphene oxide composite and their thermal stability, hydrophobic and mechanical properties, J. Asian Ceram. Soc. 5 (2017) 160-168. https://doi.org/10.1016/j.jascer.2017.04.001.
[74] F. Ciftci, S. Ayan, N. Duygulu, Y. Yilmazer, Z. Karavelioglu, M. Vehapi, R. Cakir Koc, M. Sengor, H. Yilmazer, D. Ozcimen, O. Gunduz, C.B. Ustundag, Selenium and clarithromycin loaded PLA-GO composite wound dressings by electrospinning method, Int. J. Polym. Mater. Polym. Biomater. 71 (2022) 898-909. https://doi.org/10.1080/00914037.2021.1925276.
[75] P.S.R. Dr. T. Victor Babu, S. Santosh Kumar, Design of glass fiber reinforced plastics modified with CNT and pre-stretching fabric for potential sports instruments, Mater. Des. 72 (2016) 1705-1710. https://doi.org/10.1088/1742-6596/755/1/011001.
[76] H. Belaid, S. Nagarajan, C. Teyssier, C. Barou, J. Barés, S. Balme, H. Garay, V. Huon, D. Cornu, V. Cavaillès, M. Bechelany, Development of new biocompatible 3D printed graphene oxide-based scaffolds, Mater. Sci. Eng. C. 110 (2020). https://doi.org/10.1016/j.msec.2019.110595.
[77] R.K.P. and B.C.R. D S Kumar*, M J Shukla, K K Mahato, D K Rathore, "Experimental investigation of mechanical characterization and drilling of fabricated GFRP composites reinforced with Al2O3 micro particles," in: Eff. Post-Curing Therm. Mech. Behav. GFRP Compos., 2015: p. 75.
[78] K. Kumar, P.K. Kharwar, R.K. Verma, Investigation on Flexural strength of Polymer nanocomposites modified by MWCNT and Glass fiber, E3S Web Conf. 309 (2021) 01224. https://doi.org/10.1051/e3sconf/202130901224.
[79] D.K. Singh, R.K. Verma, Development of reduced Graphene oxide modified Ultrahigh molecular weight polyethylene (rGO/UHMWPE) based Nanocomposites for Biomedical Applications, J. Thermoplast. Compos. Mater. 0 (2022) 089270572211294. https://doi.org/10.1177/08927057221129486.
[80] Y. Ajaj, H.N.K. Al-Salman, A.M. Hussein, M.K. Jamee, S. Abdullaev, A.A. Omran, M.M. Karim, A.S. Abdulwahid, Z.H. Mahmoud, others, file:///D:/UPCST-2023/PLA+G/Tensile Test/d2ra07011a.pdf file:///D:/UPCST-2023/PLA+G/Tensile Test/han2011.pdf, Case Stud. Chem. Environ. Eng. 9 (2024) 100612.
[81] S.O. Han, M. Karevan, M.A. Bhuiyan, J.H. Park, K. Kalaitzidou, Effect of exfoliated graphite nanoplatelets on the mechanical and viscoelastic properties of poly(lactic acid) biocomposites reinforced with kenaf fibers, J. Mater. Sci. 47 (2012) 3535-3543. https://doi.org/10.1007/s10853-011-6199-8.
[82] K.H. Dehnou, G.S. Norouzi, M. Majidipour, A review: studying the effect of graphene nanoparticles on mechanical, physical and thermal properties of polylactic acid polymer, RSC Adv. 13 (2023) 3976-4006. https://doi.org/10.1039/d2ra07011a.
[83] D.K.K. Cavalcanti, M.D. Banea, J.S.S. Neto, R.A.A. Lima, L.F.M. da Silva, R.J.C. Carbas, Mechanical characterization of intralaminar natural fibre-reinforced hybrid composites, Compos. Part B Eng. 175 (2019) 107149. https://doi.org/10.1016/j.compositesb.2019.107149.
[84] R.L. King, Fibre-reinforced composites materials, manufacturing and design, 1989. https://doi.org/10.1016/0010-4361(89)90651-4.
[85] P. Szatkowski, L. Czechowski, J. Gralewski, M. Szatkowska, Mechanical properties of polylactide admixed with carbon nanotubes or graphene nanopowder, Materials (Basel). 14 (2021). https://doi.org/10.3390/ma14205955.
[86] I. Taraghi, A. Fereidoon, A. Mohyeddin, The effect of MWCNTs on the mechanical properties of woven Kevlar/epoxy composites, Steel Compos. Struct. 17 (2014) 825-834. https://doi.org/10.12989/scs.2014.17.6.825.
[87] D.K. Singh, R.K. Verma, S. Mishra, Enhancement in Physiomechanical, Thermal and Morphological Characteristics of Zirconia and Reduced Graphene Oxide Modified Hybrid Polymer Nanocomposites for Orthopedic Application, J. Inorg. Organomet. Polym. Mater. (2023). https://doi.org/10.1007/s10904-023-02743-4.
[88] M. Yuan, M.J. Kummer, R.D. Milton, T. Quah, S.D. Minteer, Efficient NADH regeneration by a redox polymer-immobilized enzymatic system, ACS Catal. 9 (2019) 5486-5495.
[89] C.F. Rosewall, P.A. Sibbald, D. V. Liskin, F.E. Michael, Palladium-catalyzed carboamination of alkenes promoted by N-fluorobenzenesulfonimide via C-H activation of arenes, J. Am. Chem. Soc. 131 (2009) 9488-9489. https://doi.org/10.1021/ja9031659.
[90] J.S. Lindsey, J.N. Woodford, A Simple Method for Preparing Magnesium Porphyrins, Inorg. Chem. 34 (1995) 1063-1069. https://doi.org/10.1021/ic00109a011.
[91] A.D. Adler, F.R. Longo, F. Kampas, J. Kim, On the preparation of metalloporphyrins, J. Inorg. Nucl. Chem. 32 (1970) 2443-2445. https://doi.org/10.1016/0022-1902(70)80535-8.
,CLAIMS:1. A mechanically triggered on-demand easy-to-fabricate donor-acceptor nanocomposite photocatalyst comprising of
Nicotinamide adenine dinucleotide (ß-NAD+), formate dehydrogenase enzyme, 2,2'-bipyridine, (pentethylcyclopentadienyl) rhodium (III) chloride dimer, N, N-dimethylformamide, polylactic acid, chloroform, cyclohexane, silica gel, and magnesium chloride (MgCl2), iodine, pyrrole, dichloromethane, p-chloranil, benzaldehyde, ascorbic acid, monobasic and dibasic phosphate buffer.
2. A method for nanocomposite photocatalyst as claimed in claim 1, wherein the said method for the preparation of mechanically triggered on-demand easy-to-fabricate donor-acceptor nanocomposite photocatalyst comprising the steps of
a. cutting raw aloe vera into small pieces and washed with DI water several times after the green skin was carefully removed;
b. grounding the internal part of the aloe vera leaf in the mixer along with ethanol;
c. stirring the prepared mixture at 550 rpm along with 150°C continues until the mixed solution becomes dry;
d. transferring a specific amount of aloe vera gel to a crucible and heated in a muffle furnace at 500°C for four hours after drying;
e. removing the crucible from the muffle furnace and allowed to cool at room temperature after the heating process;
f. obtaining product aloe vera drive graphene;
g. adding benzaldehyde (1 mmol), iodine (0.21 equiv), and pyrrole (1 mmol) in turn to 10 mL of dichloromethane (DCM);
h. after adding the para-chloranil (0.75 equivalent), a second activation (300 W, 35 0C) was carried out;
i. using DCM/petroleum ether (1:1) as the eluent, the mixture was purified by flash chromatography;
j. obtaining a purple solid (59 mg, 36%) (tetra-phenyl-porphyrin);
k. allowing the solution containing 0.25 g of tetra-phenyl-porphyrin and 0.28 g of MgCl2 in 10 mL of DMF to reflux for about 1 hour;
l. allowing the solution to cool after refluxing;
m. adding 200 mL of distilled water to the cooled solution and it was maintained in an ice bath for 15 minutes;
n. following filtering, the solid precipitate of Mg-TPP was obtained; and
o. rinsing the product with distilled water and left to dry;
p. dispersing Graphene powder in chloroform (CHCl3) and then sonicated at 45 kHz for an hour;
q. adding PLA pellets along with CHCl3 and Mg-TPP to the mixture;
r. stirring the mixture at 170°C for two hours and 30 minutes;
s. pouring the resin blend into a mould sized 100 × 100 × 3 mm and dried for 24 hours under solar light;
t. cutting the fabricated composite for mechanical strength testing in accordance with ASTM standards.
3. The method as claimed in claim 2, wherein the testing mechanical strength in accordance with ASTM standards, followed by tensile, flexural, impact, and hardness tests, as well as 1,4-NADH regeneration.
4. The method as claimed in claim 2, wherein for comparative mechanical strength testing, four types of specimens were developed: P neat, 0.5, 1, and 1.5% PGT nanocomposite flexible artificial leaf photocatalyst.
5. The method as claimed in claim 2, wherein the production of formic acid from CO2 was also carried out in a quartz reactor at ambient temperature in an inert environment utilising a solar light source.
6. The method as claimed in claim 2, wherein using a solar light source and an inert atmosphere, photochemical regeneration of 1,4-NADH was carried out at ambient temperature in a quartz reactor.

Documents

NameDate
202411086442-FORM-8 [12-11-2024(online)].pdf12/11/2024
202411086442-COMPLETE SPECIFICATION [11-11-2024(online)].pdf11/11/2024
202411086442-CORRESPONDENCE-OTHERS [11-11-2024(online)].pdf11/11/2024
202411086442-FORM-9 [11-11-2024(online)].pdf11/11/2024
202411086442-DECLARATION OF INVENTORSHIP (FORM 5) [09-11-2024(online)].pdf09/11/2024
202411086442-DRAWINGS [09-11-2024(online)].pdf09/11/2024
202411086442-EDUCATIONAL INSTITUTION(S) [09-11-2024(online)].pdf09/11/2024
202411086442-EVIDENCE FOR REGISTRATION UNDER SSI [09-11-2024(online)].pdf09/11/2024
202411086442-EVIDENCE FOR REGISTRATION UNDER SSI(FORM-28) [09-11-2024(online)].pdf09/11/2024
202411086442-FORM 1 [09-11-2024(online)].pdf09/11/2024
202411086442-FORM FOR SMALL ENTITY(FORM-28) [09-11-2024(online)].pdf09/11/2024
202411086442-POWER OF AUTHORITY [09-11-2024(online)].pdf09/11/2024
202411086442-PROVISIONAL SPECIFICATION [09-11-2024(online)].pdf09/11/2024

footer-service

By continuing past this page, you agree to our Terms of Service,Cookie PolicyPrivacy Policy  and  Refund Policy  © - Uber9 Business Process Services Private Limited. All rights reserved.

Uber9 Business Process Services Private Limited, CIN - U74900TN2014PTC098414, GSTIN - 33AABCU7650C1ZM, Registered Office Address - F-97, Newry Shreya Apartments Anna Nagar East, Chennai, Tamil Nadu 600102, India.

Please note that we are a facilitating platform enabling access to reliable professionals. We are not a law firm and do not provide legal services ourselves. The information on this website is for the purpose of knowledge only and should not be relied upon as legal advice or opinion.