Home The experimental study on the air oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid with Co–Mn–Br system
Article Open Access

The experimental study on the air oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid with Co–Mn–Br system

  • Linrun Li , Suohe Yang , Haibo Jin EMAIL logo , Guangxiang He , Xiaoyan Guo and Lei Ma
Published/Copyright: November 8, 2023
Become an author with De Gruyter Brill

Abstract

2,5-Furandicarboxylic acid (FDCA), an eco-friendly biomass resource capable of replacing petroleum-based fuels, is gaining increasing popularity. In this article, 2,5-FDCA was prepared by liquid-phase oxidation of the sustainable precursor 5-hydroxymethylfurfural using the Co–Mn–Br catalyst system. The effects of catalyst concentration, catalyst ratio, reaction temperature, reaction time, reaction pressure, and solvent ratio on the reaction of FDCA were investigated. The products are subjected to qualitative and quantitative analyses using high-performance liquid chromatography, infrared spectroscopy, and hydrogen nuclear magnetic spectroscopy. Moreover, considering the loss of catalytic liquid, the suitable reaction conditions were determined as follows: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature of 170°C, reaction pressure of 2 MPa, reaction time 40 min, and airflow rate 1.0 L·min−1. Under these conditions, the yield of the product is 86.01%, the purity is 97.53%, and the loss of the catalytic liquid is about 5.63%, which is at an ideal level and provides a good basis for the recovery of the subsequent catalytic liquid and multiple cycle reactions. Through the optimization of the existing process, the use of noble metal catalysts has been reduced, and the recycling of catalytic liquid has also reduced the consumption of catalysts. This advancement marks a significant stride toward sustainable development in the green chemical industry.

1 Introduction

As the consumption of non-renewable resources continues to rise, the copious release of gases such as CO, CO2, SO2, and others has contributed to the escalation of environmental pollution and intensifies the greenhouse effect in the atmosphere, exacerbating the energy crisis. The pursuit of sustainable and eco-friendly alternatives has become an urgent priority [1]. Therefore, there is a growing preference for green biomass resources, which offer abundant content and minimal environmental pollution impact [2]. The substitution of biomass resources for fossil resources presents a viable solution to mitigate the environmental issues associated with fossil fuel combustion and alleviate the energy crisis [3].

2,5-Furandicarboxylic acid (FDCA), recognized as an essential green bio-based chemical [4,5], was included among the high-value-added bio-based chemicals by the U.S. Department of Energy in 2004 [6]. Polycondensation of FDCA with ethylene glycol (EG) enables the production of high-performance bio-based polyester materials, such as polyethylene 2,5-furan dicarboxylate (PEF). Compared with polyethylene terephthalate (PET), PEF is superior to PET in terms of tensile strength, Young’s modulus, gas barrier properties, thermal stability, and mechanical properties. Additionally, PEF offers the advantages of environmental sustainability and renewability [7,8,9]. FDCA can not only be used as the upstream monomer of new polyester materials but also be widely used in food materials, medical materials, aviation materials, and other fields [10,11,12,13,14,15,16]. Its diverse applications make it a subject of significant research interest. Therefore, it is necessary to develop sustainable, eco-friendly, and heterogeneous catalytic methods for the preparation of green biomass and its derivatives under mild reaction conditions [17].

The liquid-phase catalytic oxidation of 5-hydroxymethylfurfural (HMF) to FDCA synthesis can be divided into biosynthesis and chemical synthesis. In the biosynthesis method [18], the preparation process involving biological enzyme catalyst is complex, and the subsequent product purification is challenging. The chemical synthesis method utilizing the glycolic acid method [19] for FDCA preparation encounters issues with unfriendly reaction conditions, such as the EG rearrangement, resulting in low yields. As a consequence, its large-scale industrial applications remain limited. The furfural (furoic acid) method for FDCA production [20,21], utilizing green and environmentally friendly starting materials, faces challenges due to the presence of numerous isomers, making separation and purification difficult. Moreover, the recovery and treatment of by-product salts pose complexities, rendering it impractical. However, using HMF as raw material, the preparation of FDCA by liquid-phase oxidation is favored because of its simple process and excellent yield and purity. This process builds upon the Mid-Century liquid-phase oxidation process and demonstrates suitability for industrial production. In the single-pot operation of heterogeneous catalytic systems, the separation and purification of intermediates can be significantly eliminated and the formation of unwanted by-products can be prevented [22,23,24].

Eastman company [25] used the MC production process, which involves the liquid-phase air catalytic oxidation of furfural derivatives method such as 5-methyl-2-furfural and 5-ethyl hydroxymethyl-2-furfural, to prepare FDCA using the Co–Mn–Br catalytic system. Furanix Technologies B.V. [26] also employed the Co–Mn–Br catalytic system, but they used a mixture of one or more of these compounds as raw materials at a specific temperature and pressure. These furfural derivatives included 5-HMF, 5-methyl furfural, 5-methylfuroic acid, and 2,5-dimethylfuran. The solvent used in this process was a mixture of acetic acid and water. Canon Co. in Japan [27] has also developed a method for producing FDCA by liquid-phase catalytic oxidation of 5-HMF using the Co–Mn–Br catalytic system. At the same time, this method proposed controlling the water content in the solution to avoid catalyst deactivation and reduce combustion side reactions to improve the FDCA yield. Therefore, the Co–Mn–Br homogeneous catalytic oxidation system serves as a valuable reference and provides guidance for the large-scale preparation of FDCA.

In this study, FDCA was synthesized by liquid-phase oxidation of HMF. The effects of catalyst concentration, ratio, temperature, reaction time, pressure, and solvent ratio on the yield of oxidation reaction and the loss of reaction solvent were investigated. Through the investigation of the process conditions, the more suitable reaction conditions were obtained. Additionally, the potential for recycling the catalytic liquid was investigated, offering valuable insights for future industrial applications.

2 Experimental part

2.1 Reagents and instruments

The following reagents were used in the experiments: 5-HMF (98% purity), 2,5-FDCA (99% purity), glacial acetic acid, cobalt acetate tetrahydrate, manganese acetate tetrahydrate, hydrobromic acid (48% aqueous solution), trifluoroacetic acid, and methanol. All were obtained from Meryer (Shanghai) Chemical Technology Co. Ltd. Deionized water was used throughout the experiments.

The following experimental instruments were used: high-performance liquid chromatography (Agilent 1290), Fourier transform infrared spectrometer (Thermo Scientific Nicolet iS20), ultrasonic cleaning machine (PS-100A), electronic balance (MP4002), and high-speed centrifuge.

2.2 Experimental apparatus and methods

The experiment was carried out in a titanium reactor, as shown in Figure 1. The catalyst, raw material, and reaction solvent were mixed in the specified proportion, heated, and dissolved. The resulting mixture was then added to the titanium reactor through the feed port. Nitrogen gas was introduced to replace the air inside the reactor, while leak detection was performed simultaneously. The system pressure was adjusted, and the reaction mixture was heated and stirred. The airflow was measured using a mass flowmeter. After the completion of the reaction, the reactor was washed with glacial acetic acid to obtain a mixed solution containing FDCA. The solid–liquid separation was employed to separate the oxidation product FDCA, while the recovered catalytic liquid could be recycled for subsequent reactions.

Figure 1 
                  Reactor plant. 1. Reactor, 2. Mass flow meter, 3. One-way valve 4. Ball valve 5. Screw valve 6. Stirrer 7. Condenser 8. Tank 9. Gas cylinder 10. Feed port 11. Catalyst feed port.
Figure 1

Reactor plant. 1. Reactor, 2. Mass flow meter, 3. One-way valve 4. Ball valve 5. Screw valve 6. Stirrer 7. Condenser 8. Tank 9. Gas cylinder 10. Feed port 11. Catalyst feed port.

2.3 Product analysis method

The product was identified as FDCA by infrared spectrum analysis after drying, and the infrared spectrum of the product is shown in Figure 2. In Figure 2a, the characteristic peaks at 1,691 and 1,572 cm−1 are the product marker peaks. The characteristic absorption peak at 1,691 cm−1 corresponds to the C═O stretching vibration of the carboxylic acid group, confirming the absorption peak of carboxylic acid. The absorption peak at 1,572 cm−1 represents the C═C stretching vibration of the furan ring, further supporting the identification of FDCA. It is consistent with the Fourier transform infrared (FTIR) spectra of the standard product (as shown in Figure 2b).

Figure 2 
                  FTIR spectra of FDCA. Figure 2a is the infrared spectrum of product. Figure 2b is the infrared spectrum of standard sample.
Figure 2

FTIR spectra of FDCA. Figure 2a is the infrared spectrum of product. Figure 2b is the infrared spectrum of standard sample.

Figure 3 is the H1 NMR spectrum of the oxidation product. As shown in Figure 3a, the spectrum reveals two single peaks of H at the chemical shift δ 13.60, corresponding to H7 and H9 on the carboxyl group. Additionally, there are two single peaks of H at the chemical shift δ 7.28, representing H2 and H3 on the furan ring. Figure 3b is the H1 NMR spectra of FDCA, based on the hydrogen nuclear magnetic analysis, the structure of the oxidation product H was confirmed to be consistent with FDCA.

Figure 3 
                  H1-NMR spectra of FDCA. Figure 3a is the H1-NMR spectra of product. Figure 3b is the H1-NMR spectra of standard sample.
Figure 3

H1-NMR spectra of FDCA. Figure 3a is the H1-NMR spectra of product. Figure 3b is the H1-NMR spectra of standard sample.

3 Results and discussion

3.1 Influence of total catalyst concentration on reaction

The effects of different catalyst concentrations on the loss of oxidation products and catalytic liquid are illustrated in Figure 4. From Figure 4, with the increase of catalyst concentration, the yield and purity of the product increased first and then decreased, while the loss of catalytic liquid exhibits an upward trend. This is because, with the increase of Co–Mn–Br catalyst concentration, the active free radicals in the oxidation reaction gradually increase, accelerating the oxidation reaction and facilitating the conversion of substrate HMF to FDCA.

Figure 4 
                  Effect of catalyst concentration on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).
Figure 4

Effect of catalyst concentration on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).

However, when the catalyst concentration exceeds a certain threshold, the excessive catalytic concentration leads to intensify the condensation of intermediate products and side reactions, resulting in reduced product yield and purity. However, the increase in catalyst concentration can accelerate the initiation of free radical chain reaction, thus accelerating the oxidation reaction. However, if the concentration is too high, the free radicals are too active, which are easy to attack the furan ring, and then trigger the ring opening, so that the substrate is deeply oxidized, and the intermediate cannot be selectively converted to FDCA in time, resulting in polycondensation. At the same time, the side reaction of bromination will also be intensified, resulting in more brominated by-products, reducing the selectivity of FDCA and accelerating the corrosion of the equipment. Moreover, the increases in solvent loss occur because the solvent is susceptible to attack by active radicals or high-valent ions in the Co–Mn–Br catalytic system, triggering the dehydrogenation reaction that generates CO2 or CO and forming active radicals. Consequently, this leads to the loss of catalytic liquid. Based on the experimental results, a catalytic concentration of 6,200 PPM is deemed more suitable for this study.

3.2 Influence of n(Co + Mn)/n(HMF) on reaction

The effects of n(Co + Mn)/n(HMF) on the oxidation products and catalytic liquid loss are shown in Figure 5. From Figure 5, as the concentration of Co and Mn catalysts increases, the yield and purity of the product increase first and then decrease, while the loss of the catalytic liquid increases. The liquid-phase oxidation technology of 5-HMF is based on the liquid-phase oxidation technology of p-xylene. Both are Co–Mn–Br three-way catalytic systems with HAC as the solvent for air oxidation. There are similarities in the liquid-phase oxidation mechanism between the two processes. In the Co–Mn–Br catalytic system, Co2+ is oxidized to Co3+ by air, and Co3+ acts as a strong oxidant. Similarly, Mn2+ is oxidized to Mn3+, with the half-life of Mn3+ being much larger than that of Co3+. During the oxidation reaction, Br is oxidized to Br·, forming active free radicals and initiating chain reactions.

Figure 5 
                  Effect of n(Co + Mn)/n(HMF) on oxidation reaction. (Reaction conditions: n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).
Figure 5

Effect of n(Co + Mn)/n(HMF) on oxidation reaction. (Reaction conditions: n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).

Therefore, when the concentration of Co and Mn ions is at a low level, the formation rate of high valence metal ions is slow, which affects the whole oxidation reaction rate. As a result, incomplete reactions of the intermediate products occur, leading to the production of by-products and a decrease in yield and purity. The increase in catalytic liquid loss is due to the higher concentrations of the Co and Mn ions. When Co3+ oxidizes Mn2+, it will also make the solvent completely oxidized or deeply oxidized to produce CO2 or CO, thus increasing the solvent loss. Considering industrial production, it is advisable to avoid excessively high concentrations of Co and Mn ions. Therefore, n(Co + Mn)/n(HMF) = 0.173 is more appropriate in this experiment.

3.3 Influence of n(Br)/n(HMF) on reaction

The effects of n(Br)/n(HMF) on the loss of oxidation products and catalytic liquid are shown in Figure 6. It can be seen from Figure 6 that with the increase in Br concentration, the yield and purity of the product increased first and then decreased, while the loss of the catalytic liquid showed a downward trend. In the Co–Mn–Br catalytic system, Br plays a role in assisting catalysis. Br has a strong hydrogen absorption ability, which can capture the hydrogen on the furan ring and form active free radicals, thus triggering the chain reaction. Therefore, at the beginning of the reaction, a particular concentration of Br ions can promote the smooth progress of the chain reaction so that the substrate HMF can be deeply oxidized. However, when the concentration of Br ions is too high, it is more conducive to the occurrence of bromination side reactions, which ultimately leads to a decrease in the yield and purity of the product. The loss of the catalytic liquid gradually decreased because Co3+ was easy to form a complex with Br ions, and electron transfer occurred inside, so that Br ions replaced the acetic acid group and complexed with Co3+, thus slowing down the process of decarboxylation of the acetic acid group. Therefore, n(Br)/n(HMF) = 0.083 is more appropriate in this experiment.

Figure 6 
                  Effect of n(Br)/n(HMF) on oxidation reaction. (Reaction conditions: n(Co)/n(Mn) = 1/0.04, n(HMF)/n(HAc) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).
Figure 6

Effect of n(Br)/n(HMF) on oxidation reaction. (Reaction conditions: n(Co)/n(Mn) = 1/0.04, n(HMF)/n(HAc) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).

3.4 Influence of reaction temperature on reaction

The effects of reaction temperatures on the loss of oxidation products and catalytic liquid are shown in Figure 7. It can be seen from Figure 7 that as the reaction temperature increases, the product yield first increases and then decreases, while the product purity gradually reaches a stable state. Additionally, the loss of catalytic liquid gradually increases. The oxidation reaction is a free radical reaction, and the reaction temperature significantly influences the activity of the catalyst, which in turn affects the formation of free radicals. Gonzalez-Casamachina et al. [28] found that temperature affected the selective conversion of HMF to FDCA by affecting the formation of free radicals under visible light with O2 as oxidant and ZnO/PPy as photocatalyst. At lower reaction temperatures, the substrate HMF cannot be converted to the product in time, leading to incomplete oxidation of the intermediate and gradual accumulation, thereby impacting the yield and purity of product. On the other hand, increasing the reaction temperature can accelerate the oxidation of the substrate HMF to FDCA. However, excessively high temperatures can result in deep oxidation of the substrate, resulting in polycondensation side reactions and reduced yield. The increase in catalytic liquid loss is due to the increase in temperature, which accelerates the process of oxidative decarboxylation or decarbonylation of acetic acid. This process contributes to the loss of catalytic liquid. Based on the experimental findings, the reaction temperature of 170°C is considered more suitable for this study, as it strikes a balance between achieving a higher product yield, maintaining product purity, and minimizing catalytic liquid loss.

Figure 7 
                  Effect of reaction temperature on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).
Figure 7

Effect of reaction temperature on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).

3.5 Influence of reaction time on reaction

The effects of reaction times on the oxidation products and the loss of catalytic liquid are shown in Figure 8. It can be seen from Figure 8 that as the reaction time increases, the product yield gradually increases to a slight decrease, while the loss of catalytic liquid gradually increases. The oxidation reaction is a series of reactions that require a particular reaction time to complete oxidation. With the increased reaction time, the chain reaction can be carried out smoothly, enabling high selectivity in the oxidation of the substrate HMF.

Figure 8 
                  Effect of reaction time on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, airflow 1.0 L·min−1).
Figure 8

Effect of reaction time on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, airflow 1.0 L·min−1).

However, with the increase in reaction time, side reactions become more prevalent, leading to an increase in the possibilities of intermediate polycondensation and excessive oxidation. Consequently, this results in a reduction in the yield and purity of the product. The loss of catalytic liquid demonstrates a steady increase until the reaction time reaches 40 min. After this point, the increase in catalytic liquid loss becomes negligible. This indicates that the oxidation reaction has reached completion, and any further increase in catalytic liquid loss may be attributed to self-combustion. Based on these observations, the reaction time of 40 min is considered appropriate for this experiment. It allows for sufficient reaction time to achieve high selectivity in the oxidation process while minimizing the negative effects of side reactions on product yield and purity. Additionally, it ensures that the oxidation reaction is completed, thus minimizing the loss of catalytic liquid.

3.6 Influence of reaction pressure on reaction

The effect of reaction pressures on the oxidation products and the loss of catalytic liquid is shown in Figure 9. From Figure 9, the yield and purity of the product increased first and then decreased with the increase in reaction pressure, while the loss of catalytic liquid gradually decreased. In the liquid-phase oxidation reaction, when the pressure is low, the dissolved oxygen content in the solvent is low, and the mass transfer resistance of gas–liquid is considerable. Increasing the reaction pressure can enhance the dissolved oxygen concentration in the solvent. Enough oxygen can make the reaction substrate and the intermediate oxidation product more quickly to the substrate for high selective conversion, thus increasing the product’s purity. However, with the increase in the reaction pressure, there is also a promotion of oxidation side reactions. These side reactions can lead to undesired outcomes such as the polycondensation reaction of the oxidation intermediate, the ring opening of the furan ring, etc., which result in a decrease in the yield and purity of product. Moreover, excessive high pressure places higher demands on the material of the reactor, which can compromise safety considerations. The increase in reaction pressure makes it easier for the solvent to reach the saturated vapor pressure, slowing down the process of solvent combustion. This is also a contributing factor to the decrease in the loss of catalytic liquid. Under the premise of ensuring the safety of industrial production and maintaining the product’s yield and purity, a reaction pressure of 2.0 MPa is deemed more appropriate for this experiment.

Figure 9 
                  Effect of reaction pressure on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction time 40 min, airflow 1.0 L·min−1).
Figure 9

Effect of reaction pressure on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction time 40 min, airflow 1.0 L·min−1).

3.7 Influence of m(HMF)/m(HAC) on reaction

The effects of solvent ratios on the loss of oxidation products and catalytic liquid are shown in Figure 10. It can be seen from Figure 10 that with the decrease in solvent ratio, the yield and purity of the product increased first and then decreased, while the loss of catalytic liquid gradually decreased. The substrate HMF has high reactivity and is sensitive to reaction temperature and concentration. Under high solvent ratio conditions, HMF is more susceptible to brominated side reactions, polycondensation reactions, and other undesired reactions, which can adversely affect the yield and purity of the product. By reducing the solvent ratio, the instantaneous concentration of substrate HMF is decreased, which in turn reduces the occurrence of brominated side reactions, polycondensation reactions, and other undesired reactions.

Figure 10 
                  Effect of m(HMF)/m(HAC) on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).
Figure 10

Effect of m(HMF)/m(HAC) on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min, airflow 1.0 L·min−1).

However, when the solvent ratio becomes too low, that is, the substrate concentration is very low, a large number of free radicals attack the furan ring. This can lead to the ring-opening reaction of the furan ring and the generation of additional ring-opening by-products, resulting in a decrease in product yield and purity. Furthermore, while the lower solvent ratio can slow down the combustion of acetic acid, it may also bring about the problem of excessive solvent and increased energy consumption. Considering these factors, the solvent ratio of 1/30 is more appropriate in this experiment.

3.8 Influence of the airflow on reaction

The effect of airflow rates on the oxidation products and catalytic liquid loss is shown in Figure 11. From Figure 11, with the increase in airflow rate, the yield and purity of the product increased first and then decreased slightly, while the loss of catalytic liquid continued to increase. When the airflow rate increases, the amount of oxygen in the reaction also increases. This leads to an improvement in the solubility of oxygen in the catalytic solution and an increase in the gas–liquid-phase interface area, both of which is conducive to improving the oxidation reaction rate.

Figure 11 
                  Effect of airflow on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min).
Figure 11

Effect of airflow on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min).

However, when the airflow rate is too large, the reaction can enter an oxygen saturation state controlled by the reaction kinetics. In this state, the airflow rate has little effect on the reaction rate. Additionally, a high airflow rate reduces the residence time of air in the solvent, which can be detrimental to the reaction process. Furthermore, an increase in airflow can result in the larger amount of solvent being entrained with the exhaust gas. This not only increases the cost of exhaust gas treatment but also leads to a loss of catalytic liquid. Therefore, to ensure that the reaction is not controlled by gas phase mass transfer, it is advisable to carry out the oxidation reaction at a lower airflow rate. In this experiment, an airflow rate of 1.0 L·min−1 is more appropriate.

3.9 Influence of catalytic liquid recycling on reaction

In the first three catalytic liquid recycling, the yield and purity of the product were not significantly reduced (Figure 12), indicating that the catalyst system still had an ideal catalytic effect. However, a significant reduction in the yield and purity of the product was observed in the fifth cycle. This can be attributed to the loss of a portion of the Br catalyst during the separation of the FDCA mixture from the solid–liquid.

Figure 12 
                  Effect of catalytic liquid recycling on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min).
Figure 12

Effect of catalytic liquid recycling on oxidation reaction. (Reaction conditions: n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min).

Although the Co–Mn–Br synergistic catalytic performance is improved by adding 25% Br catalyst during the fourth catalytic liquid recycling, the accumulation of by-product water has a particular impact on the recycling of the catalytic liquid, and multiple reactions also reduce the performance of the catalytic liquid itself, resulting in a decrease in yield and purity of the product. Therefore, under suitable conditions, the catalytic liquid can be recycled multiple times, leading to a reduction in the consumption of precious metal catalysts in the reaction. This recycling capability broadens the scope of industrial application, offering economic and environmental benefits.

4 Liquid-phase oxidation reaction process of HMF

HMF oxidation follows the classical free radical oxidation mechanism and undergoes a chain reaction, generally divided into three stages: chain initiation, growth, and termination. The Co–Mn–Br system acts as a catalyst [29,30,31], facilitating peroxide decomposition and promoting the formation of free radical. The three catalytic components of Co–Mn–Br with different valence states can increase the rate of electron transfer rate within the chain reaction. These components can cycle between valence states, effectively enhancing the rate of oxidation reaction. Yang and Zhang [32,33] found that charge transfer affects the reduction rate of ions, which affects the selective conversion of substrates to products, and is the rate-determining step affecting the conversion efficiency of HMF.

During the chain initiation stage, the Co–Mn–Br system undergoes oxidation by air. Specifically, the Co2+–Br complex is oxidized to the Co3+–Br complex by air, while Co3+ oxidizes Mn2+ to Mn3+. The half-life of Mn3+ is much longer than that of Co3+, prolonging the presence of Mn3+ in the system. This leads to electron transfer from Br electron, resulting in the generation of bromine radicals. These radicals then capture hydrogen from the hydroxymethyl or aldehyde group of HMF, initiating the formation of free radicals and initiating the chain reaction. At the same time, Co2+ can be oxidized to Co3+ again. The reaction is repeated throughout the reaction, as shown in Figure 13.

Figure 13 
               Effect of catalytic liquid recycling on oxidation reaction [34].
Figure 13

Effect of catalytic liquid recycling on oxidation reaction [34].

The preparation of furan-dicarboxylic acid catalyzed by catalytic oxidation of the Co–Mn–Br system is a free radical oxidation reaction, mainly involving the oxidation of the hydroxymethyl and aldehyde group to the carboxyl group [34]. The oxidation of HMF to FDCA mainly includes two oxidation processes: alcohol oxidation to aldehyde and aldehyde oxidation to acid. The oxidation process follows the chain reaction mechanism, and the change in catalyst valence state is also coupled with each other until the end of the reaction, as shown in Figure 14.

Figure 14 
               In the process of HMF oxidation, alcohol is oxidized to aldehyde and aldehyde is oxidized to acid [34].
Figure 14

In the process of HMF oxidation, alcohol is oxidized to aldehyde and aldehyde is oxidized to acid [34].

In the process of HMF oxidation, there are three intermediate products, as shown in Figure 15, which are furan-2,5-dicarbaldehyde (DFF), 5-hydroxymethyl-2-furancarboxylic acid (HMFCA), and 5-formyl-2-furancarboxylic acid (FFCA). The selective conversion of HMF to DFF is significantly more potent than that of HMFCA, indicating that the hydroxymethyl group on the furan ring of HMF has higher activity and is more easily oxidized than the aldehyde group. It is known [35] that the C–H bond dissociation energy on the hydroxymethyl group is 309.8 kJ·mol−1, the O–H bond dissociation energy is 384.1 kJ·mol−1, and the C–H bond dissociation energy on the aldehyde group is 359.7 kJ·mol−1. The C–H bond dissociation energy on the methyl group is much lower than that of the other valence bonds, so the energy required for valence bond cleavage is the lowest, consistent with the result of high-performance liquid chromatography. In summary, this article gives the possible oxidation reaction process of HMF, as shown in Figure 16. This mechanistic understanding of the oxidation process provides insights into the key steps involved in the conversion of HMF to FDCA and the role of the Co–Mn–Br catalyst system in facilitating this transformation.

Figure 15 
               Liquid-phase spectra of FDCA. From left to right are FDCA, HMFCA, FFCA, DFF. (Above experiments were quantitatively analyzed by external standard method. The analysis conditions are as follows: the chromatographic column is a C18 reverse phase chromatographic column; the mobile phase was methanol −0.05% phosphoric acid aqueous solution (volume ratio of 55:45). The flow rate was 0.35 mL·min−1. The detector wavelength was 278 nm. The column temperature was 40°C. The injection volume was 5 μL.).
Figure 15

Liquid-phase spectra of FDCA. From left to right are FDCA, HMFCA, FFCA, DFF. (Above experiments were quantitatively analyzed by external standard method. The analysis conditions are as follows: the chromatographic column is a C18 reverse phase chromatographic column; the mobile phase was methanol −0.05% phosphoric acid aqueous solution (volume ratio of 55:45). The flow rate was 0.35 mL·min−1. The detector wavelength was 278 nm. The column temperature was 40°C. The injection volume was 5 μL.).

Figure 16 
               Liquid-phase catalytic oxidation of FDCA.
Figure 16

Liquid-phase catalytic oxidation of FDCA.

5 Conclusion

FDCA, a green biomass resource, was prepared by liquid-phase oxidation using HMF as raw material. The catalyst concentration, catalyst ratio, reaction temperature, reaction time, reaction pressure, and solvent ratio were investigated, and the loss of catalytic liquid was analyzed. The product was quantitatively and qualitatively analyzed by liquid chromatography, infrared spectroscopy, and hydrogen nuclear magnetic resonance spectroscopy.

Under the conditions of catalyst concentration 6,200 PPM, n(Co)/n(Mn)/n(Br) = 1/0.04/0.5, n(HMF)/n(HAC) = 0.05, reaction temperature 170°C, reaction pressure 2 MPa, reaction time 40 min and airflow 1.0 L·min−1, FDCA was prepared by liquid-phase oxidation. The final product yield was 86.01%, and the purity was 97.53%. Compared with the existing process [36], this method resulted in an approximately 20% increase in the product yield.

The loss of catalytic liquid under the above oxidation conditions was within a reasonable range, indicating favorable conditions for the recovery and recycling of subsequent catalytic liquid. After the catalytic liquid is recovered, multiple reaction cycles can occur. The yield and purity were not significantly reduced in the first three reaction cycles. The addition of 25% Br catalyst improved the catalytic performance, and there is a declining trend in product yield and purity. However, the activity of catalyst and the performance of catalytic liquid still decreased over time, impacting the product’s yield and purity. Therefore, the catalytic liquid can be recycled many times by adding an appropriate amount of catalyst, which expands the application range for subsequent industrialization.

The oxidation reaction of HMF exhibited a higher reactivity for the hydroxymethyl group compared to the aldehyde group. Therefore, the oxidation reaction primarily occurs in the hydroxymethyl group first. Through analyzing the oxidation reaction intermediates and considering the Co–Mn–Br catalytic mechanism, two possible paths of the HMF oxidation reaction process were speculated: HMF-DFF-FFCA-FDCA and HMF-HMFCA-FFCA-FDCA.

These findings suggest that the liquid-phase oxidation of HMF using the Co–Mn–Br catalyst system is an effective method for the production of FDCA with improved product yield and purity. The study also highlights the potential for recycling the catalytic liquid, thereby reducing the need for additional catalysts and expanding its industrial applicability.

Acknowledgments

This study was financially supported by the National Natural Science Foundation of China (22378026) and The Project of Construction of Innovative Teams and Teacher Career Development for Universities and Colleges under Beijing Municipality (IDHT20180508).

  1. Funding information: Authors state no funding involved.

  2. Author contributions: Linrun Li: writing – original draft, writing – review & editing, methodology, formal analysis; Suohe Yang: software; Haibo Jin: writing – review & editing, resources, funding acquisition, conceptualization; Guangxiang He: writing – review & editing, methodology; Xiaoyan Guo: supervision; Lei Ma: formal analysis, investigation.

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

  4. Data availability statement: The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

References

[1] Arthur J, Ragauskas AJ, Williams CK, Davison BH, Tschaplinski T. The path forward for biofuels and biomaterials. Science. 2006;311(5760):484–9.10.1126/science.1114736Search in Google Scholar PubMed

[2] Zhu Y, Romain C, Williams CK. Sustainable polymers from renewable resources. Nature. 2016;540(7633):354–62.10.1038/nature21001Search in Google Scholar PubMed

[3] Sheldon RA. Green and sustainable manufacture of chemicals from biomass: state of the art. Green Chem. 2014;16(3):950–63.10.1039/C3GC41935ESearch in Google Scholar

[4] Basil JN, Ma ADN, Libuse B, Shanks B. Plat-form biochemical for a biorenewable chemical industry. Plant J. 2008;54(4):536–45.10.1111/j.1365-313X.2008.03484.xSearch in Google Scholar PubMed

[5] Yuriy R, Juben NC, James AD. Phase modifiers promote efficient production of hydroxymethylfurfural from fructose. Science. 2006;312(5782):1933–7.10.1126/science.1126337Search in Google Scholar PubMed

[6] Bozell JJ, Petersen GR. Technology development for the production of biobased products from biorefinery carbohydrates-the US Department of Energy’s “Top 10” revisited. Green Chem. 2010;28(41):539–54.10.1002/chin.201028264Search in Google Scholar

[7] Papageorgiou GZ, Tsanaktsis V, Bikiaris DN. Synthesis of poly (ethylene furandicarboxylate) polyesterusing monomers derived from renewable resources: thermal behavior comparison with PET and PEN. Phys Chem Chem Phys. 2014;16(17):7946–58.10.1039/C4CP00518JSearch in Google Scholar PubMed

[8] Rodriguez DM. 2,5-Furandicarboxylic acid (FDCA)-a very promising building block. Melliand China. 2018;46(10):4–6.Search in Google Scholar

[9] Wang XL, Xie JS, Liao WY. Properties and Applications of Bio-based 2,5-furan Phthalate(FDCA) and Polyester(PEF). Guangdong Chem Ind. 2020;47(7):140–2.Search in Google Scholar

[10] Wu B, Xu Y, Bu Z, Wu L, Li BG, Dubois P. Biobased poly (butylene 2,5-furandicarboxylate) and poly(butyleneadipate-co-butylene 2,5-furandicarboxylate)s: From synthesis using highly purified 2,5-furandicarboxylic acid tothermo-mechanical properties. Polymer. 2014;55(16):3648–55.10.1016/j.polymer.2014.06.052Search in Google Scholar

[11] Zhou C, Gong DS, Li XM, Bai XS. Synthesis research progress and application prospects of 2,5-furandicarboxylic acid. Dyest Coloration. 2018;55(2):38–42.Search in Google Scholar

[12] Peng W. Study on Melt-spinnability, Fiber Structure and Properties of Poly(ethylene terephthalate-co-ethylene 2,5-Furandicarboxylate) copolyesters. CNKI: Ningbo University; 2018.Search in Google Scholar

[13] Lewkowski J. Synthesis, chemistry and applications of 5-hydroxymethylfurfural and its derivatives. Arkivoc. 2001;34(2):17–54.10.3998/ark.5550190.0002.102Search in Google Scholar

[14] Gandini A, Silvestre AJD, Neto CP, Sousa AF, Gomes M. The furan counterpart of poly (ethyleneterephthalate): an alternative material based on renewable resources. J Polym Sci Part A: Polym Chem. 2009;47(1):295–8.10.1002/pola.23130Search in Google Scholar

[15] Gandini A, Belgacem MN. Furan chemistry at the service of functional macromolecularmaterials: the reversible Diels-Alder reaction. Oxf Univ Press. 2007;954:280–95.10.1021/bk-2007-0954.ch018Search in Google Scholar

[16] Fang Z, Smith RL Jr, Qi XH. Production of platform chemicals fromsustainable resources. Biofuels Biorefineries. 2017;9:1659–64.10.1007/978-981-10-4172-3Search in Google Scholar

[17] Meng Y, Huang JS, Li J, Jian YM, Yang S, Li H. Enzyme-mimicking single atoms enable selectivity control in visible-light-driven oxidation/ammoxidation to afford bio-based nitriles. Green Chem. 2023;25:4453–62.10.1039/D3GC00968HSearch in Google Scholar

[18] Dijkman WP, Groothuis DE, Fraaije MW. Enzyme-catalyzed oxidation of 5-hydroxymethylfurfural to Furan-2,5-dicarboxylic acid. Angew Chem. 2014;53(25):6515–8.10.1002/anie.201402904Search in Google Scholar PubMed

[19] Li WJ. Synthesis of furan-2,5-dicarboxylic acid and its dimethyl ester catalyzed by ferrous chloride. Chem Reag. 2013;35(8):767–8.Search in Google Scholar

[20] Pan T, Deng I, Xu Q, Zuo Y, Guo QX, Fu Y. Catalytic Conversion of Furfural into a 2,5-Furandicarboxylic Acid Based Polyester with Total Carbon Utilization. Chem Sustainability Energy Mater. 2013;6(1):47–50.10.1002/cssc.201200652Search in Google Scholar PubMed

[21] Dick GR, Frankhouser AD, Banerjee A, Kanan MW. A scalable carboxylation route to furan-2,5-dicarboxylicacid. Green Chem. 2017;19(13):2966–72.10.1039/C7GC01059ASearch in Google Scholar

[22] Li H, Fang Z, Smith RLJ, Yang S. Efficient valorization of biomass to biofuels with bifunctional solid catalytic materials. Prog Energy Combust Sci. 2016;55:98–194.10.1016/j.pecs.2016.04.004Search in Google Scholar

[23] Lee JM, Na Y, Han H, Chang S. Cooperative multi-catalyst systems for one-pot organic transformations. Chem Soc Rev. 2004;33:302–12.10.1039/b309033gSearch in Google Scholar PubMed

[24] Sheldon RA. Atom efficiency and catalysis in organic synthesis. Pure Appl Chem. 2000;72:1233–46.10.1351/pac200072071233Search in Google Scholar

[25] Janka MJ, Parker KR, Shaikh AS. Oxidation process to produce a crude dry carboxylic acid product. US 8772513; 2014.Search in Google Scholar

[26] Diego Cesar MD, Matheus Adrianus D, Gerardus Johannes Maria G. Method for the preparation of 2,5-furandicarboxylic acid and for the preparation of the dialkyl ester of 2,5-furandicarboxylic acid. US 8865921B2; 2014.Search in Google Scholar

[27] Yutaka K, Miura T, Eritate S. Method of producing 2, 5-furandicarboxylic acid. US 8242292 B2; 2012-08-14.Search in Google Scholar

[28] Gonzalez-Casamachina DA, De la Rosa JR, Lucio-Ortiz CJ, Sandoval-Rangel L, García CD. Partial oxidation of 5 – hydroxymethylfurfural to 2,5-furandicarboxylic acid using O2 and a photocatalyst of a composite of ZnO/PPy under visible-light: Electrochemical characterization and kinetic analysis. Chem Eng J. 2020;393:124699.10.1016/j.cej.2020.124699Search in Google Scholar

[29] Partenheimer W. The structure of metal/bromide catalysts in acetic acid/water mixtures and its significance in autoxidation. J Mol Catal A: Chem. 2001;174:29–33.10.1016/S1381-1169(01)00169-8Search in Google Scholar

[30] Chavan SA, Halligudi SB, Srinivas D, Ratnasamy P. Formation and role of cobalt and manganese cluster complexes in the oxidation of p-Xylene. J Mol Catal A: Chem. 2000;161:49–64.10.1016/S1381-1169(00)00361-7Search in Google Scholar

[31] Xie G, Cheng YW, Li X. Catalytic mechanism of p-Xylene liquild-phase oxidation. Polyest Ind. 2002;4:1–4.Search in Google Scholar

[32] Yang ZH, Zhang BL, Yan CY, Xue ZM, Mu TC. The pivot to achieve high current density for biomass electrooxidation: Accelerating the reduction of Ni3+ to Ni2+. Appl Catal B: Environ. 2023;330:122590.10.1016/j.apcatb.2023.122590Search in Google Scholar

[33] Zhang YB, Xue ZM, Zhao XH, Zhang BL, Mu TC. Controllable and facile preparation of Co9S8–Ni3S2 heterostructures embedded with N,S,O-tri-doped carbon for electrocatalytic oxidation of 5-hydroxymethylfurfural. Green Chem. 2022;24:1721–31.10.1039/D1GC04499KSearch in Google Scholar

[34] Chen SB. Study on the liquid phase catalytic oxidation process of 5-hydroxymethylfurfural. CNKI: Zhejiang University; 2021.Search in Google Scholar

[35] Heng B, Youdi Z, Shuaibo C, Teng P, Youwei C, Lijun W, et al. Production of 2,5-furandicarboxylic acid by optimization of oxidation of 5-methyl furfural over homogeneous Co/Mn/Br catalysts. ACS Sustain Chem Eng. 2020;8(21):8011–23.10.1021/acssuschemeng.0c02574Search in Google Scholar

[36] Grushin V, Partenheimer W, Manzer LE. Oxidation of 5-(hydroxymethyl) furfural to 2,5-diformylfuran and subsequent decarbonylation to unsubstituted furan. US 20030055271; 2003.Search in Google Scholar

Received: 2023-06-29
Accepted: 2023-10-09
Published Online: 2023-11-08

© 2023 the author(s), published by De Gruyter

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

Articles in the same Issue

  1. Research Articles
  2. Value-added utilization of coal fly ash and recycled polyvinyl chloride in door or window sub-frame composites
  3. High removal efficiency of volatile phenol from coking wastewater using coal gasification slag via optimized adsorption and multi-grade batch process
  4. Evolution of surface morphology and properties of diamond films by hydrogen plasma etching
  5. Removal efficiency of dibenzofuran using CuZn-zeolitic imidazole frameworks as a catalyst and adsorbent
  6. Rapid and efficient microwave-assisted extraction of Caesalpinia sappan Linn. heartwood and subsequent synthesis of gold nanoparticles
  7. The catalytic characteristics of 2-methylnaphthalene acylation with AlCl3 immobilized on Hβ as Lewis acid catalyst
  8. Biodegradation of synthetic PVP biofilms using natural materials and nanoparticles
  9. Rutin-loaded selenium nanoparticles modulated the redox status, inflammatory, and apoptotic pathways associated with pentylenetetrazole-induced epilepsy in mice
  10. Optimization of apigenin nanoparticles prepared by planetary ball milling: In vitro and in vivo studies
  11. Synthesis and characterization of silver nanoparticles using Origanum onites leaves: Cytotoxic, apoptotic, and necrotic effects on Capan-1, L929, and Caco-2 cell lines
  12. Exergy analysis of a conceptual CO2 capture process with an amine-based DES
  13. Construction of fluorescence system of felodipine–tetracyanovinyl–2,2′-bipyridine complex
  14. Excellent photocatalytic degradation of rhodamine B over Bi2O3 supported on Zn-MOF nanocomposites under visible light
  15. Optimization-based control strategy for a large-scale polyhydroxyalkanoates production in a fed-batch bioreactor using a coupled PDE–ODE system
  16. Effectiveness of pH and amount of Artemia urumiana extract on physical, chemical, and biological attributes of UV-fabricated biogold nanoparticles
  17. Geranium leaf-mediated synthesis of silver nanoparticles and their transcriptomic effects on Candida albicans
  18. Synthesis, characterization, anticancer, anti-inflammatory activities, and docking studies of 3,5-disubstituted thiadiazine-2-thiones
  19. Synthesis and stability of phospholipid-encapsulated nano-selenium
  20. Putative anti-proliferative effect of Indian mustard (Brassica juncea) seed and its nano-formulation
  21. Enrichment of low-grade phosphorites by the selective leaching method
  22. Electrochemical analysis of the dissolution of gold in a copper–ethylenediamine–thiosulfate system
  23. Characterisation of carbonate lake sediments as a potential filler for polymer composites
  24. Evaluation of nano-selenium biofortification characteristics of alfalfa (Medicago sativa L.)
  25. Quality of oil extracted by cold press from Nigella sativa seeds incorporated with rosemary extracts and pretreated by microwaves
  26. Heteropolyacid-loaded MOF-derived mesoporous zirconia catalyst for chemical degradation of rhodamine B
  27. Recovery of critical metals from carbonatite-type mineral wastes: Geochemical modeling investigation of (bio)hydrometallurgical leaching of REEs
  28. Photocatalytic properties of ZnFe-mixed oxides synthesized via a simple route for water remediation
  29. Attenuation of di(2-ethylhexyl)phthalate-induced hepatic and renal toxicity by naringin nanoparticles in a rat model
  30. Novel in situ synthesis of quaternary core–shell metallic sulfide nanocomposites for degradation of organic dyes and hydrogen production
  31. Microfluidic steam-based synthesis of luminescent carbon quantum dots as sensing probes for nitrite detection
  32. Transformation of eggshell waste to egg white protein solution, calcium chloride dihydrate, and eggshell membrane powder
  33. Preparation of Zr-MOFs for the adsorption of doxycycline hydrochloride from wastewater
  34. Green nanoarchitectonics of the silver nanocrystal potential for treating malaria and their cytotoxic effects on the kidney Vero cell line
  35. Carbon emissions analysis of producing modified asphalt with natural asphalt
  36. An efficient and green synthesis of 2-phenylquinazolin-4(3H)-ones via t-BuONa-mediated oxidative condensation of 2-aminobenzamides and benzyl alcohols under solvent- and transition metal-free conditions
  37. Chitosan nanoparticles loaded with mesosulfuron methyl and mesosulfuron methyl + florasulam + MCPA isooctyl to manage weeds of wheat (Triticum aestivum L.)
  38. Synergism between lignite and high-sulfur petroleum coke in CO2 gasification
  39. Facile aqueous synthesis of ZnCuInS/ZnS–ZnS QDs with enhanced photoluminescence lifetime for selective detection of Cu(ii) ions
  40. Rapid synthesis of copper nanoparticles using Nepeta cataria leaves: An eco-friendly management of disease-causing vectors and bacterial pathogens
  41. Study on the photoelectrocatalytic activity of reduced TiO2 nanotube films for removal of methyl orange
  42. Development of a fuzzy logic model for the prediction of spark-ignition engine performance and emission for gasoline–ethanol blends
  43. Micro-impact-induced mechano-chemical synthesis of organic precursors from FeC/FeN and carbonates/nitrates in water and its extension to nucleobases
  44. Green synthesis of strontium-doped tin dioxide (SrSnO2) nanoparticles using the Mahonia bealei leaf extract and evaluation of their anticancer and antimicrobial activities
  45. A study on the larvicidal and adulticidal potential of Cladostepus spongiosus macroalgae and green-fabricated silver nanoparticles against mosquito vectors
  46. Catalysts based on nickel salt heteropolytungstates for selective oxidation of diphenyl sulfide
  47. Powerful antibacterial nanocomposites from Corallina officinalis-mediated nanometals and chitosan nanoparticles against fish-borne pathogens
  48. Removal behavior of Zn and alkalis from blast furnace dust in pre-reduction sinter process
  49. Environmentally friendly synthesis and computational studies of novel class of acridinedione integrated spirothiopyrrolizidines/indolizidines
  50. The mechanisms of inhibition and lubrication of clean fracturing flowback fluids in water-based drilling fluids
  51. Adsorption/desorption performance of cellulose membrane for Pb(ii)
  52. A one-pot, multicomponent tandem synthesis of fused polycyclic pyrrolo[3,2-c]quinolinone/pyrrolizino[2,3-c]quinolinone hybrid heterocycles via environmentally benign solid state melt reaction
  53. Green synthesis of silver nanoparticles using durian rind extract and optical characteristics of surface plasmon resonance-based optical sensor for the detection of hydrogen peroxide
  54. Electrochemical analysis of copper-EDTA-ammonia-gold thiosulfate dissolution system
  55. Characterization of bio-oil production by microwave pyrolysis from cashew nut shells and Cassia fistula pods
  56. Green synthesis methods and characterization of bacterial cellulose/silver nanoparticle composites
  57. Photocatalytic research performance of zinc oxide/graphite phase carbon nitride catalyst and its application in environment
  58. Effect of phytogenic iron nanoparticles on the bio-fortification of wheat varieties
  59. In vitro anti-cancer and antimicrobial effects of manganese oxide nanoparticles synthesized using the Glycyrrhiza uralensis leaf extract on breast cancer cell lines
  60. Preparation of Pd/Ce(F)-MCM-48 catalysts and their catalytic performance of n-heptane isomerization
  61. Green “one-pot” fluorescent bis-indolizine synthesis with whole-cell plant biocatalysis
  62. Silica-titania mesoporous silicas of MCM-41 type as effective catalysts and photocatalysts for selective oxidation of diphenyl sulfide by H2O2
  63. Biosynthesis of zinc oxide nanoparticles from molted feathers of Pavo cristatus and their antibiofilm and anticancer activities
  64. Clean preparation of rutile from Ti-containing mixed molten slag by CO2 oxidation
  65. Synthesis and characterization of Pluronic F-127-coated titanium dioxide nanoparticles synthesized from extracts of Atractylodes macrocephala leaf for antioxidant, antimicrobial, and anticancer properties
  66. Effect of pretreatment with alkali on the anaerobic digestion characteristics of kitchen waste and analysis of microbial diversity
  67. Ameliorated antimicrobial, antioxidant, and anticancer properties by Plectranthus vettiveroides root extract-mediated green synthesis of chitosan nanoparticles
  68. Microwave-accelerated pretreatment technique in green extraction of oil and bioactive compounds from camelina seeds: Effectiveness and characterization
  69. Studies on the extraction performance of phorate by aptamer-functionalized magnetic nanoparticles in plasma samples
  70. Investigation of structural properties and antibacterial activity of AgO nanoparticle extract from Solanum nigrum/Mentha leaf extracts by green synthesis method
  71. Green fabrication of chitosan from marine crustaceans and mushroom waste: Toward sustainable resource utilization
  72. Synthesis, characterization, and evaluation of nanoparticles of clodinofop propargyl and fenoxaprop-P-ethyl on weed control, growth, and yield of wheat (Triticum aestivum L.)
  73. The enhanced adsorption properties of phosphorus from aqueous solutions using lanthanum modified synthetic zeolites
  74. Separation of graphene oxides of different sizes by multi-layer dialysis and anti-friction and lubrication performance
  75. Visible-light-assisted base-catalyzed, one-pot synthesis of highly functionalized cinnolines
  76. The experimental study on the air oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid with Co–Mn–Br system
  77. Highly efficient removal of tetracycline and methyl violet 2B from aqueous solution using the bimetallic FeZn-ZIFs catalyst
  78. A thermo-tolerant cellulase enzyme produced by Bacillus amyloliquefaciens M7, an insight into synthesis, optimization, characterization, and bio-polishing activity
  79. Exploration of ketone derivatives of succinimide for their antidiabetic potential: In vitro and in vivo approaches
  80. Ultrasound-assisted green synthesis and in silico study of 6-(4-(butylamino)-6-(diethylamino)-1,3,5-triazin-2-yl)oxypyridazine derivatives
  81. A study of the anticancer potential of Pluronic F-127 encapsulated Fe2O3 nanoparticles derived from Berberis vulgaris extract
  82. Biogenic synthesis of silver nanoparticles using Consolida orientalis flowers: Identification, catalytic degradation, and biological effect
  83. Initial assessment of the presence of plastic waste in some coastal mangrove forests in Vietnam
  84. Adsorption synergy electrocatalytic degradation of phenol by active oxygen-containing species generated in Co-coal based cathode and graphite anode
  85. Antibacterial, antifungal, antioxidant, and cytotoxicity activities of the aqueous extract of Syzygium aromaticum-mediated synthesized novel silver nanoparticles
  86. Synthesis of a silica matrix with ZnO nanoparticles for the fabrication of a recyclable photodegradation system to eliminate methylene blue dye
  87. Natural polymer fillers instead of dye and pigments: Pumice and scoria in PDMS fluid and elastomer composites
  88. Study on the preparation of glycerylphosphorylcholine by transesterification under supported sodium methoxide
  89. Wireless network handheld terminal-based green ecological sustainable design evaluation system: Improved data communication and reduced packet loss rate
  90. The optimization of hydrogel strength from cassava starch using oxidized sucrose as a crosslinking agent
  91. Green synthesis of silver nanoparticles using Saccharum officinarum leaf extract for antiviral paint
  92. Study on the reliability of nano-silver-coated tin solder joints for flip chips
  93. Environmentally sustainable analytical quality by design aided RP-HPLC method for the estimation of brilliant blue in commercial food samples employing a green-ultrasound-assisted extraction technique
  94. Anticancer and antimicrobial potential of zinc/sodium alginate/polyethylene glycol/d-pinitol nanocomposites against osteosarcoma MG-63 cells
  95. Nanoporous carbon@CoFe2O4 nanocomposite as a green absorbent for the adsorptive removal of Hg(ii) from aqueous solutions
  96. Characterization of silver sulfide nanoparticles from actinobacterial strain (M10A62) and its toxicity against lepidopteran and dipterans insect species
  97. Phyto-fabrication and characterization of silver nanoparticles using Withania somnifera: Investigating antioxidant potential
  98. Effect of e-waste nanofillers on the mechanical, thermal, and wear properties of epoxy-blend sisal woven fiber-reinforced composites
  99. Magnesium nanohydroxide (2D brucite) as a host matrix for thymol and carvacrol: Synthesis, characterization, and inhibition of foodborne pathogens
  100. Synergistic inhibitive effect of a hybrid zinc oxide-benzalkonium chloride composite on the corrosion of carbon steel in a sulfuric acidic solution
  101. Review Articles
  102. Role and the importance of green approach in biosynthesis of nanopropolis and effectiveness of propolis in the treatment of COVID-19 pandemic
  103. Gum tragacanth-mediated synthesis of metal nanoparticles, characterization, and their applications as a bactericide, catalyst, antioxidant, and peroxidase mimic
  104. Green-processed nano-biocomposite (ZnO–TiO2): Potential candidates for biomedical applications
  105. Reaction mechanisms in microwave-assisted lignin depolymerisation in hydrogen-donating solvents
  106. Recent progress on non-noble metal catalysts for the deoxydehydration of biomass-derived oxygenates
  107. Rapid Communication
  108. Phosphorus removal by iron–carbon microelectrolysis: A new way to achieve phosphorus recovery
  109. Special Issue: Biomolecules-derived synthesis of nanomaterials for environmental and biological applications (Guest Editors: Arpita Roy and Fernanda Maria Policarpo Tonelli)
  110. Biomolecules-derived synthesis of nanomaterials for environmental and biological applications
  111. Nano-encapsulated tanshinone IIA in PLGA-PEG-COOH inhibits apoptosis and inflammation in cerebral ischemia/reperfusion injury
  112. Green fabrication of silver nanoparticles using Melia azedarach ripened fruit extract, their characterization, and biological properties
  113. Green-synthesized nanoparticles and their therapeutic applications: A review
  114. Antioxidant, antibacterial, and cytotoxicity potential of synthesized silver nanoparticles from the Cassia alata leaf aqueous extract
  115. Green synthesis of silver nanoparticles using Callisia fragrans leaf extract and its anticancer activity against MCF-7, HepG2, KB, LU-1, and MKN-7 cell lines
  116. Algae-based green AgNPs, AuNPs, and FeNPs as potential nanoremediators
  117. Green synthesis of Kickxia elatine-induced silver nanoparticles and their role as anti-acetylcholinesterase in the treatment of Alzheimer’s disease
  118. Phytocrystallization of silver nanoparticles using Cassia alata flower extract for effective control of fungal skin pathogens
  119. Antibacterial wound dressing with hydrogel from chitosan and polyvinyl alcohol from the red cabbage extract loaded with silver nanoparticles
  120. Leveraging of mycogenic copper oxide nanostructures for disease management of Alternaria blight of Brassica juncea
  121. Nanoscale molecular reactions in microbiological medicines in modern medical applications
  122. Synthesis and characterization of ZnO/β-cyclodextrin/nicotinic acid nanocomposite and its biological and environmental application
  123. Green synthesis of silver nanoparticles via Taxus wallichiana Zucc. plant-derived Taxol: Novel utilization as anticancer, antioxidation, anti-inflammation, and antiurolithic potential
  124. Recyclability and catalytic characteristics of copper oxide nanoparticles derived from bougainvillea plant flower extract for biomedical application
  125. Phytofabrication, characterization, and evaluation of novel bioinspired selenium–iron (Se–Fe) nanocomposites using Allium sativum extract for bio-potential applications
  126. Erratum
  127. Erratum to “Synthesis, characterization, and evaluation of nanoparticles of clodinofop propargyl and fenoxaprop-P-ethyl on weed control, growth, and yield of wheat (Triticum aestivum L.)”
Downloaded on 13.11.2025 from https://www.degruyterbrill.com/document/doi/10.1515/gps-2023-0116/html?lang=en&srsltid=AfmBOoqIEbO5_UIMGAKkTFd7HjjaOUnGMvDfU2SZbzED_tGxPg83gh-x
Scroll to top button