Next Article in Journal
Computational Analysis of Structure–Activity Relationships in Highly Active Homogeneous Ruthenium−Based Water Oxidation Catalysts
Next Article in Special Issue
Controllable Construction of IrCo Nanoclusters and the Performance for Water Electrolysis
Previous Article in Journal
Facile Synthesis and Environmental Applications of Noble Metal-Based Catalytic Membrane Reactors
Previous Article in Special Issue
Insights into the Design of An Enzyme Free Sustainable Sensing Platform for Efavirenz
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Electrocatalytic Application of Copper- and Cobalt-Carbon Composites Based on Pyrolyzed Polymer

by
Zainulla M. Muldakhmetov
,
Nina M. Ivanova
*,
Yakha A. Vissurkhanova
and
Yelena A. Soboleva
Institute of Organic Synthesis and Chemistry of Coal of Kazakhstan Republic, Alikhanov Str., 1, Karaganda 100000, Kazakhstan
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(8), 862; https://doi.org/10.3390/catal12080862
Submission received: 29 June 2022 / Revised: 25 July 2022 / Accepted: 1 August 2022 / Published: 5 August 2022

Abstract

:
Copper- and cobalt-containing carbon composites were prepared by pyrolysis of an aniline-formaldehyde polymer (AFP) doped with the metal oxides, followed by the reduction of metal cations in an electrochemical cell. AFP + metal oxide nanocomposites were synthesized by introducing a metal salt during the polycondensation of aniline with formaldehyde and by alkaline precipitation of metal oxides into the polymer matrix. The heat treatment was carried out at 400, 500 and 700 °C. Microscopic studies revealed the formation of CuO crystallites in the shape of "stars" on the heat-treated carbon material. The resulting composites were saturated with hydrogen in an electrochemical system, which was accompanied by the reduction of copper and cobalt cations, and the appearance of the metals in zero-valence state. The so-prepared Cu + copper oxides/C and Co + Co(OH)2/C composites were used as electrocatalysts in the electrohydrogenation of acetophenone (APh). Compared to the electrochemical reduction of APh on a copper cathode (without catalysts), an increase in the rate of this process (by 2–4 times) in the presence of the composites and an increase in the APh conversion with the selective formation of 1-phenylethanol were established.

1. Introduction

Metal–carbon nanocomposites as multifunctional materials attract especially close attention as catalysts from metal (or metals) nanoparticles deposited on carbon supports or incorporated into the carbon matrix [1,2,3,4]. With the discovery of new carbon materials, such as carbon nanotubes, fullerenes, graphene, etc., new opportunities have appeared for producing efficient and selective catalysts and electrocatalysts. Due to their large specific surface area, good stability, and high electrical conductivity, they are not only efficient substrates, but also stabilizers for metal nanoparticles.
A common way of producing metal-carbon nanocomposites is the pyrolysis (or carbonization) of various organic materials, including natural (cellulose, rice husks, nut shells, etc.), synthetic organic compounds with both small molecules and polymers in the presence of metal-containing compounds, and the thermolysis of organometallic complexes [5,6,7]. In the recently proposed method (polymerized-complex method) [8], monometallic and bimetallic alloy nanoparticles deposited on a carbon support were prepared through the formation of a polymer–metal complex followed by its calcination in a nitrogen atmosphere. Many investigations have been devoted to the study of the various physicochemical properties of metal nanoparticles deposited on carbon doped with various heteroatoms, a review of which can be found in [9]. Of special interest are the composites of metal nanoparticles deposited on N-doped carbon [10,11,12,13], in which an increase in catalytic activity occurs due to the interaction of the metal with the nitrogen atom in the carbon matrix. Melamine and melamine-formaldehyde polymers [14,15,16,17], dicyandiamide [18], polyaniline [19], polyacrylonitrile [20,21], etc., are often used as sources of nitrogen, or nitrogen and carbon, when creating metal/N-C composites.
This work describes the fabrication of metal- and metal oxide-containing carbon composites synthesized by introducing Cu2+ and Co2+ metal cations in the in situ polycondensation of aniline with formaldehyde and the formation of metal oxides + aniline-formaldehyde polymer (AFP) composites and their further heat treatment at temperatures of 400, 500 and 700 °C. The composites obtained after the carbonization process were saturated with hydrogen in an electrochemical system, and then, in their presence, the electrocatalytic hydrogenation of acetophenone (APh) was carried out to test their electrocatalytic activity.
Acetophenone is a typical representative of aromatic ketones and often an object of research in the reduction processes using electrochemical [22,23], electrocatalytic [24,25] and catalytic [26,27,28,29,30,31] methods including newly developed catalytic systems. The electrocatalytic and catalytic hydrogenation of APh on Pt, Pd, and Ru catalysts is carried out with the formation of not only the products of hydrogenation of the carbonyl group, but also of the phenol ring. According to the article data, on such catalysts as Ni@C, Ni-B-P/SiO2, Ag-OMS-2 (silver-incorporated octahedral molecular sieve catalyst), Co/Mordenite, Cu-Zn-Al, the APh hydrogenation occurs selectively with the formation of the main product, 1-phenylethanol (1-PhE), widely employed in the pharmaceutical and fragrance industries.
Note that the AF polymer, which has nitrogen atoms in its structure, is practically not used to create N-doped carbon materials, since its heat treatment is accompanied by high loss of nitrogen atoms. Thus, the study [32] describes the preparation of nanostructured composites based on the AF polymer with introduced nickel hydroxide and subsequent carbonization at 550–850 °C in an atmosphere of pure nitrogen. The absence of nitrogen in the carbonized samples of the AF polymer and its complete decomposition at 850 °C were established. The Ni/C composites were then modified and studied for the effect of microwave radiation on them. The synthesis of cobalt nanoparticles deposited on carbon using the AF polymer was a subject of a recently published study [33]. The copper-polymer composite on the basis of AFP was also synthesized and studied for antimicrobial activity [34].
In the paper, copper- and cobalt-containing carbon composites prepared by the pyrolysis of aniline-formaldehyde polymer + metal oxides precursors and used as electrocatalysts in the APh electrohydrogenation are very active and selective to produce 1-PhE.

2. Results and Discussion

Heat treatment of the AF polymer promotes its curing and the acquisition of a cross-linked structure; its loose structure is visible in the micrograph of Figure 1a.
For the thermally treated AF polymer samples, their specific surface area was determined by the BET method (Brunauer–Emmett–Teller method) using nitrogen adsorption–desorption isotherms. Table 1 shows the values of the specific surface area (Ssp.) of the studied samples, the sizes and the total volume of pores in them (VƩ pores). It follows from these data that with an increase in the temperature of HT, the value of the specific surface area of the black carbon based on the AF polymer decreases. At the same time, the pore size distribution changes unevenly: in the AFP sample (400 °C), mesopores with sizes from 2 to 50 nm predominate, in the AFP sample (500 °C) there are only macropores with sizes larger than 50 nm, and in the AFP sample (700 °C) there are also those and others with high content of macropores (Table 1).
According to the procedure for the synthesis of the AF polymer, it is precipitated from the reaction solution by introducing sodium hydroxide. In the presence of a metal salt in the reaction medium of polycondensation, with the addition of sodium hydroxide, the metal is also co-precipitated in the form of its hydroxide, followed by decomposition to an oxide upon drying at 80 °C. The required amount of NaOH for precipitating the metal introduced in the salt form was taken into account during the syntheses performance. It can be noted that with an increase in the HT temperature, the weight of the AFP composites decreased due to a decrease in the carbon phase and, accordingly, the metal/carbon ratios changed too. The TGA results (Figure 2) for the AFP and its composites with the metal oxides after the synthesis show that the polymer itself was stable up to 400 °C, and the subsequent mass loss occurred up to 700 °C. The decomposition of the polymer base in both metal-polymer composites began at lower temperatures (at 300 °C for Cu- and 200 °C for Co-containing composite) and finished around 500 °C. This means that the presence of metal oxides in the composites reduces the thermal stability of their polymer base.
From X-ray phase analysis it follows that the phase constitution of the AFP + Cu(NO3)2 composite before heat treatment contains crystalline phases of copper (II) oxide and the polymer amorphous phase (Figure 3a). During heat treatment, copper oxide CuO was partially reduced to Cu2O oxide, and the content of this oxide was higher in the composite sample formed at 400 °C (Figure 3b) than in the samples treated at 500 and 700 °C (Figure 3c,d). These samples also contained impurities of sodium chloride, which was poorly washed out from the AF polymer base. That is, the copper-containing AFP composites after heat treatment are taking the CuO + Cu2O/C constitution with different ratios of copper oxides (Figure 3). The electrochemical saturation of the composites formed after HT with hydrogen was accompanied by the partial reduction of copper cations from its oxides, and crystalline phases of copper Cu0 appeared in their constitutions (Figure 3b–d). In this case, the content of Cu2O oxide increases with a simultaneous decrease in the content of CuO. From the XRD patterns (Figure 3b–d) it follows that the copper content in the composites increases with growing heat treatment temperature. After saturation with hydrogen, these composites have the Cu0 + Cu2O (+CuO)/C constitution.
During the heat treatment of AFP + Co(NO3)2 composite, the crystalline phases of cobalt oxides (Co3O4, CoO) and an amorphous phase of AFP soot with NaCl admixture were formed (Figure 4a). After electrochemical experiments in a catholyte alkaline solution, the crystalline phases of cobalt hydroxide (β-Co(OH)2) and reduced metal (Co0) appeared in this composite (Figure 4b).
Micrographs of the AFP + Cu(NO3)2 composite after HT (Figure 5) depict the changes in the carbonized AF-polymer with an increase in HT temperature and show the distribution of copper oxide particles inside and over the surface of the polymer-carbon base. Thus, the porous structure of the AFP carbon base becomes denser with an increase in the HT temperature, which corresponds to the data on the determination of the specific surface area of the carbonized polymer (Table 1). Copper oxide particles were localized on the AFP carbon support surface in small and larger collections. The sizes of these particles were ~50–90 nm according to microscopic measurements.
In the sample treated at 700 °C, crystallites of copper oxide CuO were found in the form of four-pointed "stars" located with a weak degree of order on the surface of the carbonized AF polymer (Figure 5c). At a larger scale of magnification, one can see that these “stars” also had a fifth branch directed upwards. The “stars” consist of the thin plates with a diameter of ~135–160 nm self-organizing into such unusual structures. The nitrogen content in the composition of the carbon matrix decreased sharply with an increase in the HT temperature, and in composites treated at 700 °C energy-dispersive X-ray spectroscopy (EDS X-ray) analyses showed its absence.
The micrographs of the AFP + Co(NO3)2 (700 °C) composite after heat treatment (Figure 6a) show an almost uniform distribution of cobalt oxide particles over the surface and inside the polymer-carbon base; their size was ~50–100 nm. In the same composite after its application in the electrohydrogenation of acetophenone, the rather large crystallites in the form of hexagonal disks (or prisms) of various heights were detected (Figure 6b). These disks are assembled from the thinnest plates (~25–30 nm), which can be seen at higher magnification. According to EDS X-ray analysis, they consist of elements such as Co and O, but the H element may also be present, which is not determined by such analyses. Since it follows from the XRD analysis that the composite sample contains mainly the β-Co(OH)2 crystalline phases and a smaller amount of reduced cobalt, it can be assumed that the hexagonal disks are formed from cobalt hydroxide. In the literature, e.g., [35], there is a description of hexagonal crystals in the form of plates prepared using chemical methods and consisting of both cobalt hydroxide and cobalt oxides. In our studies, such plates (or disks) were formed under the conditions of electrochemical experiments. It should be noted that the Co + Co(OH)2/C composite produced after the electrochemical experiments had magnetic properties obviously due to the presence of reduced cobalt.
The composites synthesized on the basis of AF polymer were saturated with hydrogen in an electrochemical cell with the possibility of undergoing the reduction of metal cations during this period, and then they were used to activate the cathode in the electrohydrogenation of acetophenone. The results of the electrocatalytic hydrogenation of APh on the thermally treated AFP composite samples are displayed in Table 2. The experiments were carried out in two different cells (cell 1 and cell 2), which are described in Section 3.
It follows from the data in Table 2 that thermally treated Cu-containing composites absorb hydrogen in the electrochemical system, and the volume of absorbed hydrogen (VH2) increases with an increase in HT temperature of the samples. The content of reduced copper in these composites increased in the same dependence (Figure 3b–d). Among the Co-containing composites, H2 absorption was observed only for the composite with HT at 700 °C (Table 2, cell 1). The reduction of cobalt cations occurred in a small amount under the given conditions and apparently only from CoO.
The electrochemical reduction of APh on a Cu cathode (without a catalyst) was carried out with its low conversion and the formation of dimeric products (pinacones, PCs) in a fairly large amount (cell 1, Table 2). The APh electrocatalytic hydrogenation carried out using Cu- and Co-containing carbon composites as electrocatalysts intensified this process. The Cu + Cu2O (+ CuO)/C composites prepared on the base of AFP + Cu(NO3)2 samples demonstrated the highest catalytic activity: the rate of APh hydrogenation increased by 2–3 times (in cell 1) and by 4 times (in cell 2) compared to its electrochemical reduction on the Cu cathode. Moreover, its increase was directly related to the temperature of heat treatment of the prepared polymer-metal oxide composites: the higher it was, the higher the rate of APh hydrogenation. Referring to the XRD analysis (Figure 3), this is due to an increase in the content of electrochemically reduced copper in the composite samples. It should be noted that the use of Cu- and Co-containing AFP-carbon composites as electrocatalysts in the process significantly increased the selectivity of the 1-phenylethanol formation compared with the APh electrochemical reduction (Table 2).
Additionally, the electrocatalytic activity of metal-oxide carbon composites heat-treated at 500 °C was compared with their precursors, polymer-metal oxide composites without HT (in Table 2, these are AFP + Cu(NO3)2 and AFP + Co(NO3)2 composites). Experiments were performed in cell 2. The results showed that, first, AFP composites without HT also exhibited electrocatalytic activity in the electrohydrogenation of APh with a high conversion and selective formation of 1-PhE compared to the electrochemical reduction of APh on the Cu cathode. Second, the rate of APh hydrogenation on these composites was slightly decreased compared to this process in the presence of composites with HT at 500 °C. Perhaps such a decrease in the hydrogenation rate is due to their lower content of copper and cobalt (Table 2). It follows from these experiments that polymer–metal oxide composites, in which metal cations are able to reduce in an electrochemical cell, can serve as hydrogenation catalysts in the electrosynthesis of organic compounds even without thermal treatment, which was also shown in our earlier studies [36,37].
The durability of the resulting composites as catalysts in the process under study was tested using the AFP + Cu(NO3)2 (500 °C) sample. In the presence of the composite, three experiments on the APh electrocatalytic hydrogenation were carried out in cell 2 (with an insignificant loss of mass of this composite in each step). The results showed a slight decrease in the rate of APh hydrogenation, while its conversion remained the same high with the selective formation of 1-PhE (Table 2).
According to the literature data [31], the selective formation of 1-PhE is due to the interaction of catalyst metals with the π-electron system of the carbonyl group of APh, and not with the phenol ring. Obviously, in our studies, the reduced Cu and Co metals also interacted with the C=O group facilitating its hydrogenation under the given conditions.

3. Materials and Methods

3.1. Materials

Aniline for the synthesis of the aniline-formaldehyde polymer and acetophenone were preliminarily subjected to distillation. The remaining reagents were used as purchased: Cu(NO3)2·3H2O, Co(NO3)2·6H2O, formalin, hydrochloric acid, sodium hydroxide. All these reagents were available from “Ridder” LLP (Karaganda, Kazakhstan), except for HCl—it is from “Karagandareaktivsbyt” LLP (Karaganda, Kazakhstan). Distilled water was used to prepare aqueous solutions.

3.2. Synthesis of the Metal Oxide-AFP Composites

In the typical synthesis experiment [36], aniline (0.1 M) was mixed with 56 mL distilled water in an ice bath. After 30 min, 100 mL of 2M HCl was added to the solution and stirred for 1 h. The mixture was heated to 40 °C. Next, 30 mL of an aqueous solution of metal nitrate (Cu(NO3)2·3H2O, Co(NO3)2·6H2O), in which the initial amount of metal was 3.60 g, was added dropwise to aniline hydrochloride solution. Stirring was continued for 1 h at 40 °C. Further, 0.15 M of formalin (in the ratio of 1.5:1 to aniline) was added dropwise to this mixture and stirred for 1 h. The AFP composite was isolated from the reaction mixture by the adding a calculated amount of 25% NaOH alkali solution to it, and stirred for 1 h for homogenization. The resulting precipitate was then filtered and washed with distilled water heated to 50 °C followed by drying the precipitate in an oven at 80 °C to constant weight. The resulting powder was thoroughly ground.
The addition of sodium hydroxide to the reaction mixture leads not only to the AF-polymer separation, but also to the precipitation of metal hydroxides to the polymer matrix. Therefore, the required amounts of NaOH for the complete precipitation of copper (II) and cobalt (II) hydroxides were calculated using the following reaction equation (where M = Cu, Co): M(NO3)2 + 2NaOH = M(OH)2↓ + 2NaNO3.
The metal contents in the filtrates after the isolation of the polymer–metal composites were checked by complexometric titration using EDTA sodium salt and the murexide indicator. The conducted experiments established the absence of copper and cobalt cations in the analyzed filtrates.

3.3. Heat Treatment of the Metal Oxide-AFP Composites

The metal oxide-containing AFP-composite was divided into 3 equal parts and each part was thermally treated in the corundum crucible with a closed lid at 400, 500 and 700 °C for 2 h. As a result, metal oxides/AF-carbon composites were obtained.

3.4. Electrochemical Experiments

Powders of carbonized samples of AFP composites (weighing 1 g) were deposited on a horizontally located copper cathode (with an area of 0.05 and 0.126 dm2 in cell 1 and cell 2, respectively) tightly adjacent to the bottom of the cathode part in diaphragm electrochemical cells. The saturation of these composites with hydrogen was carried out at a current of 1.5 A (cell 1) and 2.5 A (cell 2) (current densities were 30.0 and 19.8 A/dm2) and a temperature of 30 °C in an aqueous alkaline solution of the catholyte. A Pt grid served as the anode.
Under the same conditions, after the termination of hydrogen absorption, the electrocatalytic hydrogenation of APh was carried out in the presence of formed Cu + Cu2O (+CuO)/C and Co + Co(OH)2/C composites as catalysts. Acetophenone was preliminarily dissolved in ethanol and then added to the catholyte. The initial concentrations of APh in the catholyte were 0.198 (cell 1) and 0.119 M (cell 2). The hydrogenation products were extracted from the catholyte with chloroform, and the resulting extract was analyzed using a Kristallux-4000M chromatograph.

3.5. Physical-Chemical Investigations

The specific surface area and pore size of thermally treated AF polymer samples (without a metal salt) were determined by the BET method (Brunauer–Emmett–Teller) using nitrogen adsorption–desorption isotherms on a Sorbi MS instrument (META, Moscow, Russia).
A thermogravimetric analysis (TGA) of AFP and its composites with metal oxides, which were obtained after the synthesis, was performed over a temperature range of 30–800 °C with a heating rate of 10 °C/min under air atmosphere on a LabSYS evo TGA/DTA/DSC analyzer (SETARAM Instrumentation, Caluire, France).
The changes in the phase constitution of the initially synthesized AFP composites, metal oxide/carbon composites after heat treatment, and the same composites after electrochemical experiments were controlled on a D8 ADVANCE ECO X-ray diffractometer (Bruker, Karlsruhe, Germany) using CuKα radiation in the angle range (2θ) of 15–90°. The morphological features of metal-carbon composites after the heat treatment and after electrohydrogenation of an organic compound were studied using a TESCAN MIRA 3 LMU scanning electron microscope (TESCAN, Brno, Czech Republic).

4. Conclusions

First, the aniline-formaldehyde polymer + metal oxide composites were prepared by introducing copper(II) and cobalt(II) nitrates into the polycondensation reaction of aniline with formaldehyde. Second, their heat treatment at 400, 500, and 700 °C was accompanied by the formation of metal oxide/carbon composites. In the AFP + Cu(NO3)2 (700 °C) composite, the microscopic studies revealed the formation of self-organizing structures from CuO in the shape of "stars". Third, in the electrochemical system, copper cations and, to a lesser extent, cobalt cations were reduced, and the Cu + Cu2O (+CuO)/C and Co + Co(OH)2/C composites were formed. The occurrence of hexagonal plates from cobalt hydroxide proceeded during the electrochemical experiments. The electrocatalytic properties of the prepared composites were studied in the electrohydrogenation of acetophenone. In the presence of these composites, the hydrogenation rate increased, the acetophenone conversion with selective formation of the main product, 1-phenylethanol, was also enhanced compared with the electrochemical reduction of acetophenone on the Cu cathode. Thus, the developed method for the preparation of metal-carbon composites can be used to create similar composites based on the various polymer and metal salts for their application as catalysts and electrocatalysts.

Author Contributions

Conceptualization: N.M.I.; methodology: N.M.I., Y.A.V.; investigation: Y.A.V., Y.A.S.; resources: Y.A.S.; writing—original draft preparation and editing: N.M.I.; supervision, Z.M.M.; project administration: Z.M.M.; funding acquisition: N.M.I., Z.M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Education and Science of the Republic of Kazakhstan, Grant number AP08855930.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rodrigues-Reinoso, F. The role of carbon materials in heterogeneous catalysis. Carbon 1998, 36, 159–164. [Google Scholar] [CrossRef]
  2. Wu, G.; Santandreu, A.; Kellogg, W.; Gupta, S.; Ogoke, O.; Zhang, H.; Wang, H.-L.; Dai, L. Carbon nanocomposite catalysts for oxygen reduction and evolution reactions: From nitrogen doping to transition-metal addition. Nano Energy 2016, 29, 83–110. [Google Scholar] [CrossRef] [Green Version]
  3. Gerber, I.C.; Serp, P. A theory/experience description of support effects in carbon-supported catalysts. Chem. Rev. 2020, 120, 1250–1349. [Google Scholar] [CrossRef] [PubMed]
  4. Lei, C.; Lyu, S.; Si, J.; Yang, B.; Li, Z.; Lei, L.; Wen, Z.; Wu, G.; Hou, Y. Nanostructures carbon based heterogeneous electrocatalysts for oxygen evolution reaction in alkaline media. ChemCatChem 2019, 11, 5855–5874. [Google Scholar] [CrossRef]
  5. Pomogailo, A.D.; Rozenberg, A.S.; Dzhardimalieva, G.I. Termolysis of metallopolymers and their precursors as a method for the preparation of nanocomposites. Russ. Chem. Rev. 2011, 80, 257–292. [Google Scholar] [CrossRef]
  6. Pavlenko, V.V.; Anurov, S.A.; Mansurov, Z.A.; Biisenbayev, M.A.; Azat, S.; Tanirbergenova, S.K. Thermo-oxidative modification of plant fiber. News Nat. Acad. Sci. Repub. Kazakhstan Ser. Chem. Technol. 2014, 5, 47–52. [Google Scholar]
  7. Wu, Z.-Y.; Xu, S.-L.; Yan, Q.-Q.; Chen, Z.-Q.; Ding, Y.-W.; Li, C.; Liang, H.-W.; Yu, S.-H. Transition metal-assisted carbonization of small organic molecules toward functional carbon materials. Sci. Adv. 2018, 4, eaat0788. [Google Scholar] [CrossRef] [Green Version]
  8. Li, P.; Li, F.; Deng, G.; Guo, X.; Liu, H.; Jiang, W.; Wang, T. Polymerized complex method for preparation of supported bimetallic alloy and monometallic nanoparticles. Chem. Comm. 2016, 52, 2996–2999. [Google Scholar] [CrossRef]
  9. Rangraz, Y.; Heravi, V.V.; Elhampour, A. Recent advances in heteroatom-doped porous carbon/metal materials: Fascinating heterogeneous catalysts for organic transformations. Chem. Rec. 2021, 21, 1985–2073. [Google Scholar] [CrossRef]
  10. Mabena, L.F.; Ray, S.S.; Mhlanga, S.D.; Coville, N.J. Nitrogen-doped carbon nanotubes as a metal catalyst support. Appl. Nanosci. 2011, 1, 67–77. [Google Scholar] [CrossRef] [Green Version]
  11. Shen, W.Z.; Fan, W.B. Nitrogen-containing porous carbons: Synthesis and application. J. Mater. Chem. A 2013, 1, 999–1013. [Google Scholar] [CrossRef]
  12. Majeed, S.; Zhao, J.; Zhang, L.; Anjum, S.; Liu, Z.; Xu, G. Synthesis and electrochemical applications of nitrogen-doped carbon nanomaterials. Nanotechnol. Rev. 2013, 2, 615–635. [Google Scholar] [CrossRef]
  13. Thakur, A.K.; Kurtyka, K.; Majumder, M.; Yang, X.; Ta, H.Q.; Bachmatiuk, A.; Liu, L.; Tizebicka, B.; Rummeli, M.H. Recent advances in boron- and nitrogen-doped carbon-based materials and their various applications. Adv. Mater. Interfaces 2022, 9, 2101964. [Google Scholar] [CrossRef]
  14. Pretschuh, C.; Schwarzinger, C.; Abdala, A.A.; Vukusic, S. Characterization of conductive nanographite melamine composites. Open J. Compos. Mater. 2014, 4, 61–71. [Google Scholar] [CrossRef] [Green Version]
  15. Tiwari, D.; Goel, C.; Bhunia, H.; Bajpai, P.K. Melamine-formaldehyde derived porous carbons for adsoption of CO2 capture. J. Environ. Manag. 2017, 197, 415–427. [Google Scholar] [CrossRef]
  16. Zhong, H.; Zhang, H.; Liu, S.; Deng, C.; Wang, M. Nitrogen-enriched carbon from melamine resins with superior oxygen reduction reaction activity. ChemSusChem 2013, 6, 807–812. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, X.; Huang, Y.; Chen, X.; Gao, Q.; Zhang, W. Functionalized metal-organic frameworks for effective removal of rocephin in aqueous solutions. J. Colloid Interface Sci. 2018, 514, 234–239. [Google Scholar]
  18. Choi, C.H.; Park, S.H.; Woo, S.I. N-doped carbon prepared by pyrolysis of dicyandiamide with various MeCl2·xH2O (Me = Co, Fe, and Ni) composites: Effect of type and amount of metal seed on oxygen reduction reactions. Appl. Catal. B Environ. 2012, 119–120, 123–131. [Google Scholar] [CrossRef]
  19. Daems, N.; Wouters, J.; Van Goethem, C.; Baert, K.; Poleunis, C.; Delcorte, A.; Hubin, A.; Vankelecom, I.F.J.; Pescarmona, P.P. Selective reduction of nitrobenzene to aniline over electrocatalysts based on nitrogen-doped carbons containing non-noble metals. Appl. Catal. B Environ. 2018, 226, 509–522. [Google Scholar] [CrossRef]
  20. Kozhitov, L.V.; Muratov, D.G.; Yakushko, E.V.; Kozhitov, S.L.; Savchenko, A.G.; Shchetinin, I.V.; Emelyanov, S.G.; Chervjakov, L.M. The synthesis of metalcarbon nanocomposite Ni/C on the basis of polyacrylonitrile. J. Nano- Electron. Phys. 2013, 5, 5855–5874. [Google Scholar]
  21. Zaporotskova, I.A.; Kozhitov, L.V.; Anikeev, N.A.; Davletova, O.A.; Popkova, A.V.; Muratov, D.G.; Yakushko, E.V. Metal-carbon nanocomposites based on pyrolysed polyacrylonitrile. Mod. Electron. Mater. 2015, 1, 43–49. [Google Scholar] [CrossRef]
  22. Szikra, D.; Dublinszky, K.; Cserhati, C.; Nagy, I.P. Elaboration of composite electrodes based on Raney Nickel and Pd/C for electrosynthesis application. J. New Mater. Electrochem. Syst. 2010, 13, 133–140. [Google Scholar]
  23. Bondue, C.J.; Koper, M.T. Electrochemical reduction of the carbonyl functional group: The importance of adsorption geometry, molecular structure, and electrode surface structure. J. Am. Chem. Soc. 2019, 141, 12071–12078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Vilar, M.; Oliveiro, J.L.; Navarro, M. Investigation of the hydrogenation reactivity of some organic substrates using an electrocatalytic method. Appl. Catal. A Gen. 2010, 372, 1–7. [Google Scholar] [CrossRef]
  25. Villalba, M.; del Pozo, M.; Calvo, E.J. Electrocatalytic hydrogenation of acetophenone and benzophenone using palladium electrodes. Electrochem. Acta. 2015, 164, 125–131. [Google Scholar] [CrossRef]
  26. Yadav, G.D.; Mewada, R.K. Selective hydrogenation of acetophenone to 1-phenyl ethanol over nanofibrous Ag-OMS-2 catalysts. Catal. Today 2012, 198, 330–337. [Google Scholar] [CrossRef]
  27. Lin, S.; Liu, J.; Ma, L. Ni@C catalyzed hydrogenation of acetophenone to phenyl ethanol under industrial mild conditions in a flow reactor. React. Chem. Eng. 2022, 7, 1126–1135. [Google Scholar] [CrossRef]
  28. Wang, J.; Wang, Y.; Chen, G.; He, Z. Highly loaded and dispersed Ni2P/Al2O3 catalyst with high selectivity for hydrogenation of acetophenone. Catalysts 2018, 8, 309. [Google Scholar] [CrossRef] [Green Version]
  29. Zhang, X.B.; Lu, W.N.; Wang, J.L.; Liao, B.H.; Qin, Y.H.; Zhang, Y.J.; Zhang, B.; Xin, S.A. Selective catalytic hydrogenation of acetophenone to 1-phenylethanol over Co/mordenite in water. Catal. Commun. 2019, 119, 124–128. [Google Scholar] [CrossRef]
  30. Pei, F.; Wang, Z.; Lv, J.; Li, F.; Xue, W. Efficient catalytic transfer hydrogenation of acetophenone to 1-phenylethanol over Cu-Zn-Al catalysts. Ind. Eng. Chem. Res. 2022, 61, 5419–5428. [Google Scholar] [CrossRef]
  31. Bertero, N.M.; Apesteguia, C.R.; Marchi, A.J. Catalytic and kinetic study of the liquid-phase hydrogenation of acetophenone over Cu/SiO2 catalyst. Appl. Catal. A Gen. 2008, 349, 100–109. [Google Scholar] [CrossRef]
  32. Saini, L.; Patra, M.K.; Dhaka, M.K.; Jani, R.K.; Gupta, G.K.; Dixit, A.; Vadera, S.R. Ni/Graphitic core-shell nanostructured based light weight elastomeric composites for Ku-band microwave absorbing applications. Cryst. Eng. Comm. 2018, 20, 4630–4640. [Google Scholar] [CrossRef]
  33. Chaudhary, J.; Tailor, G.; Verma, D.; Verma, R. Synthesis and characterization of cobalt nanocomposite using aniline-formaldehyde resin. Compos. Comm. 2020, 18, 13–18. [Google Scholar] [CrossRef]
  34. Chaudhary, J.; Tailor, G.; Tomer, N. Synthesis, characterization and studies on antimicrobial properties of copper nanocomposite. Mat. Sci. Ind. J. 2018, 16, 130. [Google Scholar]
  35. Yang, J.; Liu, H.; Martens, W.N.; Frost, R.L. Synthesis and characterization of cobalt hydroxide, cobalt oxyhydroxide, and cobalt oxide nanodiscs. J. Phys. Chem. C 2010, 114, 111–119. [Google Scholar] [CrossRef]
  36. Ivanova, N.M.; Visurkhanova, Y.A.; Soboleva, E.A.; Pavlenko, N.A.; Muldakhmetov, Z.M. Structure and electrocatalytic activity of aniline-formaldehyde polymer doped with copper(II) chloride. ChemistrySelect 2016, 1, 5304–5309. [Google Scholar] [CrossRef]
  37. Ivanova, N.M.; Soboleva, E.A.; Visurkhanova, Y.A.; Lazareva, E.S. Structure-phase changes in polymer composites doped with silver nitrate and their electrocatalytic activity. Russ. J. Electrochem. 2018, 54, 1010–1017. [Google Scholar] [CrossRef]
Figure 1. (a) SEM image of the AF polymer heat treated at 200 °C; (b) its proposed structure.
Figure 1. (a) SEM image of the AF polymer heat treated at 200 °C; (b) its proposed structure.
Catalysts 12 00862 g001
Figure 2. TGA curves of AFP (red color), AFP + Cu(NO3)2 (blue) and AFP + Co(NO3)2 (green).
Figure 2. TGA curves of AFP (red color), AFP + Cu(NO3)2 (blue) and AFP + Co(NO3)2 (green).
Catalysts 12 00862 g002
Figure 3. XRD patterns of AFP + Cu(NO3)2 composites (a) before HT and heat treated at (b) 400 °C, (c) 500 °C and (d) 700 °C.
Figure 3. XRD patterns of AFP + Cu(NO3)2 composites (a) before HT and heat treated at (b) 400 °C, (c) 500 °C and (d) 700 °C.
Catalysts 12 00862 g003
Figure 4. XRD patterns of AFP + Co(NO3)2 composites (a) after the HT at 700 °C and (b) after APh electrohydrogenation.
Figure 4. XRD patterns of AFP + Co(NO3)2 composites (a) after the HT at 700 °C and (b) after APh electrohydrogenation.
Catalysts 12 00862 g004
Figure 5. SEM images of AFP + Cu(NO3)2 composite samples heat treated at 400 °C (a), 500 °C (b) и 700 °C (c).
Figure 5. SEM images of AFP + Cu(NO3)2 composite samples heat treated at 400 °C (a), 500 °C (b) и 700 °C (c).
Catalysts 12 00862 g005
Figure 6. SEM images of AFP + Co(NO3)2 (700 °C) composite (a) after HT and (b) after APh electrohydrogenation.
Figure 6. SEM images of AFP + Co(NO3)2 (700 °C) composite (a) after HT and (b) after APh electrohydrogenation.
Catalysts 12 00862 g006
Table 1. Specific surface area and porous structure of AFP samples after heat treatment.
Table 1. Specific surface area and porous structure of AFP samples after heat treatment.
Polymer
after HT
Ssp.,
m2/g
VƩ Pores,
cm3/g
Pore Size,
nm
Pore Size
Distribution, %
AFP (400 °C)44.8 ± 23.70.4393.5
29.4
56.1
79.6
22.7
35.4
20.6
21.3
AFP (500 °C)23.0 ± 2.70.16156.1
79.6
56.2
43.8
AFP (700 °C)13.9 ± 0.60.3485.9
56.1
79.6
30.5
59.6
9.9
Table 2. The results of acetophenone electrocatalytic hydrogenation using Cu- and Co-containing AFP-carbon composites.
Table 2. The results of acetophenone electrocatalytic hydrogenation using Cu- and Co-containing AFP-carbon composites.
CompositesMetal Content in 1 g of Composites,
g
VH2,1 mLElectrocatalytic
Hydrogenation of APh
Product
Compositions, %
W,2 mL H2/minα,3 %1-PhEAPhPCs
Cu cathode (cell 1)-0.02.028.14.449.945.7
AFP + Cu(NO3)2 (400 °C)
AFP + Cu(NO3)2 (500 °C)
AFP + Cu(NO3)2 (700 °C)
AFP + Cu(NO3)2 (800 °C)
0.355
0.408
0.509
0.600
13.0
43.3
60.7
106.8
4.4
4.8
5.2
6.3
89.0
94.3
95.1
89.7
95.1
96.5
97.6
96.9
4.0
2.8
2.4
3.1
0.9
0.7
-
-
Cu cathode (cell 2)-0.01.715.012.983.63.5
AFP + Cu(NO3)2
AFP + Cu(NO3)2 (500 °C)
AFP + Cu(NO3)2 (500 °C)
AFP + Cu(NO3)2 (500 °C)
0.234
0.408
<0.408
<0.408
67.3
46.0
0.0
0.0
7.7
8.8
8.4
7.4
99.5
99.6
99.8
99.6
98.5
99.0
99.6
99.2
1.2
1.0
0.4
0.5
0.3
-
-
0.3
Cu cathode (cell 1)-0.02.028.14.449.945.7
AFP + Co(NO3)2 (400 °C)
AFP + Co(NO3)2 (500 °C)
AFP + Co(NO3)2 (700 °C)
0.238
0.285
0.357
0.0
0.0
29.4
3.5
3.8
5.1
80.8
69.8
85.4
90.4
88.8
93.9
9.3
11.2
6.1
0.2
-
-
Cu cathode (cell 2)-0.01.715.012.983.63.5
AFP + Co(NO3)2
AFP + Co(NO3)2 (500 °C)
0.171
0.285
41.8
0.0
4.2
7.3
99.7
99.8
97.5
97.9
2.5
2.1
-
-
1 The volume of hydrogen absorbed by the AFP composites; 2 the average rate of APh hydrogenation over the period of its conversion α = 25%, 3 the APh conversion.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Muldakhmetov, Z.M.; Ivanova, N.M.; Vissurkhanova, Y.A.; Soboleva, Y.A. Preparation and Electrocatalytic Application of Copper- and Cobalt-Carbon Composites Based on Pyrolyzed Polymer. Catalysts 2022, 12, 862. https://doi.org/10.3390/catal12080862

AMA Style

Muldakhmetov ZM, Ivanova NM, Vissurkhanova YA, Soboleva YA. Preparation and Electrocatalytic Application of Copper- and Cobalt-Carbon Composites Based on Pyrolyzed Polymer. Catalysts. 2022; 12(8):862. https://doi.org/10.3390/catal12080862

Chicago/Turabian Style

Muldakhmetov, Zainulla M., Nina M. Ivanova, Yakha A. Vissurkhanova, and Yelena A. Soboleva. 2022. "Preparation and Electrocatalytic Application of Copper- and Cobalt-Carbon Composites Based on Pyrolyzed Polymer" Catalysts 12, no. 8: 862. https://doi.org/10.3390/catal12080862

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

Article Metrics

Back to TopTop