Next Article in Journal
Applications of Deep Eutectic Solvents in Sample Preparation and Extraction of Organic Molecules
Next Article in Special Issue
Comparing the Colloidal Stabilities of Commercial and Biogenic Iron Oxide Nanoparticles That Have Potential In Vitro/In Vivo Applications
Previous Article in Journal
Nutritional and Therapeutic Properties of Fermented Camel Milk Fortified with Red Chenopodium quinoa Flour on Hypercholesterolemia Rats
Previous Article in Special Issue
Plasmid-DNA Delivery by Covalently Functionalized PEI-SPIONs as a Potential ‘Magnetofection’ Agent
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unexpected Room Temperature Ferromagnetism of a Ball-Milled Graphene Oxide—Melamine Mixture

by
Vladimir P. Vasiliev
1,
Eugene N. Kabachkov
1,
Alexander V. Kulikov
1,
Roman A. Manzhos
1,
Iurii G. Morozov
2 and
Yury M. Shulga
1,3,*
1
Federal Research Center of Problems of Chemical Physics and Medicinal Chemistry of RAS, Acad. Semenov Ave., 1, 142432 Chernogolovka, Russia
2
Merzhanov Institute of Structural Macrokinetics and Materials Science of RAS, Acad. Osipyan St., 8, 142432 Chernogolovka, Russia
3
Department of Functional Polymer Materials, National University of Science and Technology MISIS, Leninsky Ave., 4, 119049 Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(22), 7698; https://doi.org/10.3390/molecules27227698
Submission received: 28 September 2022 / Revised: 15 October 2022 / Accepted: 6 November 2022 / Published: 9 November 2022
(This article belongs to the Special Issue Advanced Research in Magnetic Nanoparticles)

Abstract

:
Nitrogen-doped carbon nanomaterial (NDCNM) was synthesized by grinding a mixture of graphene oxide and melamine in a planetary mill with both balls and milling chamber of zirconium dioxide. In the electron spin resonance spectrum of NDCNM, a broad signal with g = 2.08 was observed in addition to a narrow signal at g = 2.0034. In the study using a vibrating-sample magnetometer, the synthesized material is presumably a ferromagnet with a coercive force of 100 Oe. The specific magnetization at 10,000 Oe is approximately 0.020 and 0.055 emu/g at room temperature and liquid nitrogen temperature, respectively.

1. Introduction

Organic ferromagnetism was first loudly announced in Ref. [1]. It reported the ferromagnetism of the product of polymerization of paramagnetic stable biradical, 1,4-bis-(2,2,6,6-tetramethyl-4-oxy-4-piperidyl-1-oxyl)-butadiin, (BIPO) and described previous experimental and theoretical work on the designated topics. In particular, the theoretical prediction of 1D organic ferromagnets was noted to be proposed by Ovchinnikov in 1978 [2].
The next bright page in this story is the work of [3], which was published in 2001. The authors wrote that as a result of C60 polymerization under high pressure and high temperature, a material is formed that exhibits features characteristic of ferromagnets: “saturation magnetization, large hysteresis and attachment to a magnet at room temperature”. However, 5 years later the authors published a communication in which they doubt their results [4]. This dramatic publication is not often cited. Yet, we believe that the problem raised by this offers very instructive experience.
At present, quite a lot of publications are devoted to the ferromagnetism of d0 materials (see, for example, [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20]). There are also many works about the ferromagnetism of graphene-like structures [21,22,23,24,25,26,27]. The situation with carbon-based materials seems to us to be very interesting. For example, ferromagnetism has been observed in pure carbon structures such as the already mentioned rhombohedral C60 fullerite [3], highly oriented pyrolytic graphite [28,29], glassy carbon after laser ablation [30], and double-layer graphene [31]. Ferromagnetism was discovered for hydrogenated C60 fullerite [32] and graphene [33]. There are also numerous publications where ferromagnetism has been found in graphene oxide, reduced graphene oxide, and heteroatom doped GO and/or rGO [34,35,36,37,38,39,40,41,42,43,44,45,46,47,48]. One might assume that the issue was resolved unambiguously in favor of the existence of ferromagnetic organic and carbon materials. However, there are experimental works which prove that paramagnetic defects in graphene structures, even at their high concentration, are still far enough apart for an (anti)ferromagnetic exchange to occur between them. Of particular note is the work of Geim et al., who studied the effect of fluorination and irradiation with hydrogen and carbon atoms on the magnetic properties of graphene [49]. Based on the results of their study, the authors concluded that defective ferromagnetism is impossible in graphene structures. The absence of ferromagnetic or superparamagnetic contribution to the magnetization of graphene oxide measured up to 5 K was also stated in [50].
A simple method of production of an efficient platinum-free oxygen reduction reaction electrocatalyst was described in [51,52]. The electrocatalyst was synthesized by a solid-phase method as a result of grinding graphene oxide and melamine in a planetary ball-mill machine without the use of any solvents or high-temperature processing. Based on the XPS and IR spectroscopy data, high electrocatalytic activity of the obtained material was assumed to be determined by the presence of nitrogen atoms and quinone groups on its surface.
In this study we present the results of the analysis of obtained material by electron paramagnetic resonance (ESR) and magnetometry methods. Surprisingly, the ESR spectrum contains a fairly intense broad signal with a g factor of 2.08 in addition to the narrow ESR signal with a g factor of 2.0034, which is characteristic of materials based on graphene oxide. A vibrating-sample magnetometry study showed that the resulting material is presumably a ferromagnet with a specific saturation magnetization of 0.02 emu/g at room temperature. Elemental analysis, IR and X-ray photoelectron spectroscopy were used to characterize the samples.

2. Experimental

2.1. Synthesis of Nitrogen-Doped Carbon Material

Graphene oxide was synthesized using a modified Hummers’ method [53]. Melamine, C3N6H6 (99.9%, BASF SE (Ludwigshafen am Rhein, Germany)) was used as a source of nitrogen.
The mechanochemical synthesis was carried out in a «FRITSCH pulverisette 6» planetary mill. Grinding vessel and balls were of ZrO2, the internal diameter of the grinding vessel, the volume, and the ball diameter were 65 mm, 85 mL, and 10 mm, respectively. The GO/melamine ratio was 4:1, rotation speed was 400 rpm and grinding time was 6 min. After grinding, the resulting powder was kept for 1 h in a 10% aqueous solution of ammonia, treated in an ultrasonic bath, and then it was centrifuged and washed with water 4–5 times to remove melamine residues [52].

2.2. Characterization

Elemental analysis of the samples preliminarily degassed in an argon flow at a temperature of 80 °C for 30 min was carried out on a Vario Micro cube CHNS analyzer (Elementar GmbH, Hanau, Germany).
The IR spectra were recorded at room temperature in the range of 400–4000 cm−1 on a Perkin-Elmer “Spectrum Two” Fourier-transform spectrometer (Waltham, MA, USA) with an ATR attachment with a diamond crystal.
The ESR spectra of the powders were recorded with a Bruker Elexsys II E 500 ESR spectrometer and an SE/X 2544 radio spectrometer (Radiopan, Poznan, Poland) at room temperature. The number of spins N and the g factor were determined using the Xepr software. To check the correctness of these procedures, a weighed sample of CuSO4·5H2O and DPPH with a g factor of 2.0036 were used. The accuracy of concentration determination was ca. 15%.
Magnetic characteristics were measured using an M4500 vibrating-sample magnetometer (EG&G PARC, Gaithersburg, MD, USA) in magnetic fields up to 10 kOe at room temperature and liquid nitrogen temperature. The diamagnetic signal of the nylon sample holder was subtracted from the obtained total magnetization of material.
XPS spectra were obtained using a Specs PHOIBOS 150 MCD electronic spectrometer for chemical analysis (SPECS GmbH, Berlin, Germany). During the measurement of spectra, the vacuum in the spectrometer chamber did not exceed 2 × 10−10 Torr; the X-ray tube was equipped with a magnesium anode (Mg Kα radiation is 1253.6 eV) and the source power was 225 W. The survey spectrum was recorded in the range of 0–1000 eV in the constant transmission energy mode (40 eV for the survey spectrum and 10 eV for individual lines). The survey spectrum was recorded with a step of 1.0 eV, while the spectra of individual lines with a step of 0.05 eV.

3. Results

3.1. Elemental Analysis and SEM

Table 1 presents the data on the composition of graphene oxide (before and after grinding), melamine (before and after grinding), and grinding products of two GO:melamine mixtures after removal of unreacted melamine. As is seen, the grinding does not very strongly affect pure GO and melamine. However, the composition of a grinding product of the GO:melamine mixture is not a simple sum of the initial components. Doping of graphene oxide with nitrogen can be assumed to occur as a result of grinding. Since the product mainly contains carbon, hereinafter we will refer to it as nitrogen-doped carbon nanomaterial (NDCNM). We suggest that presence of sulfur in samples is due to the GO production by Hummers’ method. The intense washing of graphite oxide by water, probably, is not sufficient to completely remove traces of sulfuric acid, which might be trapped in closed pores (see [54]).
The SEM micrographs of graphene oxides are well known. The GO micrograph obtained by us (Figure 1) resembles those given in [55,56,57]. The surface of the GO sample consists of smooth sheets that have curves and folds. One can also distinguish the edges of individual sheets. The surface of the NDCNM sample is formed of rounded particles, the dimensions of which are generally less than 100 nm. We failed to notice the morphological features typical for GO sheets in the NDCNM micrograph.

3.2. IR Spectra

IR spectra of graphene oxide, melamine, and NDCNM—the product of mechanochemical treatment of their 4:1 mixture—is presented in Figure 2. Comparing the spectra, one can note that the IR spectrum of NDCNM differs from the spectra of initial components. In the NDCNM spectrum, there are no absorption bands of stretching vibrations of N–H bonds, which can be observed in the spectrum of pure melamine powder at 3468, 3417, 3324 and 3121 cm−1 [58]. The absence of these peaks can be interpreted as dissociation of N–H bonds as well as the complete removal of amine groups as a result of grinding. At the same time, the appearance of a number of absorption bands in the NDCNM spectrum in the region from 2350 to 1900 cm−1 can be ascribed to the presence of cyano groups in various configurations in the sample [59,60]. Note, IR spectra were obtained in vacuum, so we exclude the manifestation of vibrations of CO2 molecules of the gaseous phase in this region. Finally, in our opinion, the wide absorption band in the region from 3700 to 3000 cm−1 in the GO and NDCNM spectra is mainly due to the stretching vibrations of O–H groups bound by hydrogen bonds, which determines such a wide absorption band [61]. However, the full width at half maximum of this band in the NDCNM spectrum is less than that in the GO spectrum, and this band is shifted towards higher wavenumbers (3333 cm−1 for NDCNM and 3173 cm−1 for GO). This means that hydrogen bonds in NDCNM are weaker than the hydrogen bonds in graphene oxide. It is well known that strengthening of hydrogen bonds shifts the frequency of O−H vibrations to low frequencies by hundreds of reciprocal centimeters [62].
Comparing the spectrum of NDCNM with the spectrum of GO, one can see that the absorption band at 1730 cm−1 due to the stretching vibrations of C=O bonds [63] is practically absent in the NDCNM spectrum.

3.3. XPS Spectra

The elemental composition of some samples was also calculated using analytical lines of the survey XPS spectrum (Figure 3). As is seen, the content of oxygen in the layer analyzed by XPS is substantially less for NDCNM than for GO sample. The carbon content in the NDCNM is slightly higher (by 2.4 at.%) than that in initial GO; that is, graphene oxide is slightly reduced as a result of grinding (Table 2). As expected, the content of nitrogen in NDCNM sample is rather high. It should be mentioned that even during long-term data acquisition, no peaks of the iron group elements (Fe, Co, Ni) appeared in the XPS spectra.
The identification of surface nitrogen-containing groups can be carried out on the basis of an analysis of the fine structure of the N1s line in the spectrum. According to the literature (see [64] and references therein), pyridinic nitrogen (N1) appears in the range of 398.0–399.3 eV and pyrrolic nitrogen (N2) appears in the range of 399.8–401.2 eV) in the XPS spectra of nitrogen-doped carbon materials. Of note, the N1s lines of amino (399.1 eV [64]) and cyano (399.3 eV [64]) groups are also located in this region. Therefore, it is difficult to clearly identify them. The peak corresponding to the nitrogen atoms of the N4 type (inside a graphite sheet) is located at about 401 eV, and the peak corresponding to terminal graphite nitrogen (N3) is at 402.3 eV. Oxidized pyridine nitrogen (N5) corresponds to a peak at 402.8 eV. Physically adsorbed nitrogen (N5) corresponds to a peak at 404.7 eV [64].
Three peaks can be distinguished in the N1s spectrum of NDCNM (Figure 4a). Unexpectedly, the major contribution to the N1s line can be assigned to the pyrrolic nitrogen (N2).
C1s spectrum of NDCNM (Figure 4b) is also deconvoluted into four peaks at 284.7 (peak 1), 286.8 (peak 2), 288.7 eV (peak 3) and 290.6 eV (peak 4). These peaks can be assigned to carbon atoms not bonded with oxygen atoms (peak 1), carbon atoms singly bonded with an oxygen atom (peak 2), and carbon atoms which have two bonds with one or two oxygen atoms (peak 3) [65]. According to [66], the concentration of C(sp3) in NDCNM is 30–35%.

3.4. ESR Spectra

ESR spectra of initial GO and GO ground in a planetary mill are presented in Figure 5. As is seen, one narrow signal is observed in the ESR spectra of the samples at room temperature. At the same time, a narrow peak with g = 2.0034 is present in the NDCNM spectrum along with a broad line with g = 2.08. The nature of the narrow ESR signal for graphene oxide samples has been studied in many works. At present, this signal is considered to be due to paramagnetic centers associated with isolated carbon atoms in highly functionalized regions of graphene oxide, as well as to delocalized π electrons of aromatic domains [67]. The signal from paramagnetic centers of the first type corresponds to g > 2.003 and a half-width of 2.5 Oe; the signal from π electrons is wider (up to 12 Oe) with g = 2.002. The obtained ESR spectra of initial and ground GO indicate that localized paramagnetic centers predominate in the sample of initial GO, while milling causes a shift towards delocalized states, which is primarily evidenced by an increase in the signal half-width. In the case of NDCNM, the half-width of the narrow signal is also greater than that for initial GO. The nature of the broad signal has not yet been described in the literature.

3.5. Magnetic Measurements

σ(H) curves were measured at two temperatures (78 and 296 K) with a change in the magnetic field from −10 to 10 kOe at a constant rate of 1.2 kOe/min. The NDCNM sample showed σ(H) dependence with a well-resolved hysteresis loop typical of a ferromagnetic sample (Figure 6). The coercive force was about 100 Oe. Within the experimental error, as the temperature was increased from 78 to 296 K, the magnetization at 10 kOe decreased from ca. 0.05 to ca. 0.02 emu/g. This indicates that the Curie temperature is slightly above 300 K.

4. Discussion

In each experimental work dealing with ferromagnets with a low saturation magnetization, it is necessary to keep in mind the previous findings, which make it possible to separate impurity ferromagnetism from internal ferromagnetism. Foremost, for experimental work it is necessary to analyze the studied materials for the traces of the iron group elements (Fe, Co, Ni). Moreover, to estimate the level of impurity ferromagnetism, it is necessary to use the values of saturation magnetization σs equal to 2.2, 1.7, and 0.6 μB/atom for Fe, Co, and Ni, respectively [32]. It is desirable to check all initial materials for the presence of ferromagnetic impurities to avoid the use of magnetic stirrers, iron tweezers and spatulas and rerun experiments with unexpected results. Note that the sources of ferromagnetic impurities in the study of lightweight samples with a weak magnetic signal by such methods as vibrating-sample magnetometers (VSM) and superconducting quantum interference devices (SQUID) are well described in [8].
As for the data provided in the present study, it seems to us that impurity ferromagnetism cannot explain the obtained results. The materials were ground in a mill, all parts of which were made of non-magnetic zirconium dioxide. Moreover, grinding the graphene oxide and melamine separately did not lead to the formation of ferromagnetic substances. Ferromagnetic structures were obtained only in the case of milling a mixture of graphene oxide and melamine. One of the ferromagnetic samples was carefully analyzed by XPS and no impurities of iron group elements were found in it. The synthesis was repeated four times with slight changes in the procedure of washing the ground sample from unreacted melamine. The results on magnetic properties were reproduced with high accuracy.
Potassium permanganate is known to be often used in the oxidation of graphite [53]. The removal of manganese residues from graphene oxide is also known to be a rather laborious operation. In principle, the ferromagnetism of some manganese compounds with perovskite structure is known (see, for example, [68]). However, it is difficult to imagine that the presence of melamine contributes to the formation of a perovskite structure. We believe that the presence of a small impurity of manganese cannot explain the ferromagnetism observed by us in the product of milling a mixture of graphene oxide and melamine.

5. Conclusions

In this research, we report a simple method for producing nitrogen-doped carbon nanomaterial (NDCNM). Elemental analysis, IR, XPS, EPR, and magnetometry were used to characterize the material. Furthermore, along with the narrow ESR signal characteristic of materials based on graphene oxide, an intense broad signal with a g factor of 2.08 was detected in the ESR spectrum of the NDCNM sample. A study using a vibrating-sample magnetometer showed that, in terms of its magnetic behavior, the NDCNM can be considered with a high degree of probability to be a weak ferromagnet with a coercive force of 100 Oe and specific magnetization at 10,000 Oe, approximately 0.020 and 0.055 emu/g at room temperature and liquid nitrogen temperature, respectively.
Summarizing the available literature data on this topic, we can conclude that a huge amount of experimental work has been carried out to identify ferromagnetism in various carbon materials. Various theoretical models have been constructed to explain this phenomenon. However, despite this, the appearance of ferromagnetism in carbon materials can still be considered unexpected.

Author Contributions

Conceptualization, Y.M.S.; Data curation, V.P.V. and R.A.M.; Formal analysis, V.P.V.; Funding acquisition, I.G.M.; Investigation, V.P.V., E.N.K., A.V.K. and I.G.M.; Resources, E.N.K.; Software, E.N.K.; Supervision, A.V.K.; Validation, Y.M.S.; Visualization, R.A.M.; Writing—original draft, I.G.M.; Writing—review & editing, Y.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The study was performed in accordance with the State Assignments AAAA-A19-119032690060-9 and AAAA-A19-119061890019-5. Y.M.S. thanks the state program of basic research “for the long-term development and ensuring the competitiveness of society and the state” (47 GP) based on the universities, project number 0718-2020-0036. The work has been performed using the equipment of the FRC PCP MC RAS Research Center.

Conflicts of Interest

The authors declare that they have no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Korshak, Y.V.; Medvedeva, T.V.; Ovchinnikov, A.A.; Spector, V.N. Organic polymer ferromagnet. Nature 1987, 326, 370–372. [Google Scholar] [CrossRef]
  2. Ovchinnikov, A.A. Multiplicity of the ground state of large alternant organic molecules with conjugated bonds. Theor. Chim. Acta 1978, 47, 297–304. [Google Scholar] [CrossRef]
  3. Makarova, T.L.; Sundqvist, B.; Höhne, R.; Esquinazi, P.; Kopelevich, Y.; Scharff, P.; Davydov, V.A.; Kashevarova, L.S.; Rakhmanina, A.V. Magnetic carbon. Nature 2001, 413, 716–718. [Google Scholar] [CrossRef] [PubMed]
  4. Makarova, T.L.; Sundqvist, B.; Höhne, R.; Esquinazi, P.; Kopelevich, Y.; Scharff, P.; Davydov, V.; Kashevarova, L.S.; Rakhmanina, A.V. Retraction Note: Magnetic carbon. Nature 2006, 440, 707. [Google Scholar] [CrossRef] [Green Version]
  5. Coey, J.M.D. d0 ferromagnetism. Solid State Sci. 2005, 7, 660–667. [Google Scholar] [CrossRef]
  6. Coey, J.M.D.; Stamenov, P.; Gunning, R.D.; Venkatesan, M.; Paul, K. Ferromagnetism in defect-ridden oxides and related materials. New J. Phys. 2010, 12, 053025. [Google Scholar] [CrossRef]
  7. Ghosh, S.; Khan, G.G.; Mandal, K. d0 ferromagnetism in oxide nanowires: Role of intrinsic defects. EPJ Web Conf. 2013, 40, 03001. [Google Scholar] [CrossRef] [Green Version]
  8. Qi, B.; Ólafsson, S.; Gíslason, H.P. Vacancy defect-induced d0 ferromagnetism in undoped ZnO nanostructures: Controversial origin and challenges. Prog. Mater. Sci. 2017, 90, 45–74. [Google Scholar] [CrossRef]
  9. Garg, S.; Gautam, S.; Singh, J.P.; Kandasami, A.; Goya, N. Characterizing the defects and ferromagnetism in metal oxides: The case of magnesium oxide. Mater. Charact. 2021, 179, 111366. [Google Scholar] [CrossRef]
  10. Chouhan, L.; Bouzerar, G.; Srivastava, S.K. d0 ferromagnetism in Li-doped ZnO compounds. J. Mater. Sci. Mater. Electron. 2021, 32, 6389–6397. [Google Scholar] [CrossRef]
  11. Chouhan, L.; Srivastava, S.K. A comprehensive review on recent advancements in d0 ferromagnetic oxide materials. Mater. Sci. Semicond. Process. 2022, 147, 106768. [Google Scholar] [CrossRef]
  12. Chouhan, L.; Bouzerar, G.; Srivastava, S.K. Effect of Mg-doping in tailoring d0 ferromagnetism of rutile TiO2 compounds for spintronics application. J. Mater. Sci. Mater. Electron. 2021, 32, 11193–11201. [Google Scholar] [CrossRef]
  13. Rani, N.; Chahal, S.; Kumar, P.; Kumar, A.; Shukla, R.; Singh, S.K. MgO nanostructures at different annealing temperatures for d0 ferromagnetism. Vacuum 2020, 179, 109539. [Google Scholar] [CrossRef]
  14. Chouhan, L.; Narzary, R.; Dey, B.; Panda, S.K.; Manglam, M.K.; Roy, L.; Brahma, R.; Mondal, A.; Kar, M.; Ravi, S.; et al. Tailoring room temperature d0 ferromagnetism, dielectric, optical, and transport properties in Ag-doped rutile TiO2 compounds for spintronics applications. J. Mater. Sci. Mater. Electron. 2021, 32, 28163–28175. [Google Scholar] [CrossRef]
  15. Chouhan, L.; Srivastava, S.K. Observation of room temperature d0 ferromagnetism, band-gap widening, zero dielectric loss and conductivity enhancement in Mg doped TiO2 (rutile + anatase) compounds for spintronics applications. J. Solid State Chem. 2022, 307, 122828. [Google Scholar] [CrossRef]
  16. Chouhan, L.; Panda, S.K.; Bhattacharjee, S.; Das, B.; Mondal, A.; Parida, B.N.; Brahma, R.; Manglam, M.K.; Kar, M.; Bouzerar, G.; et al. Room temperature d0 ferromagnetism, zero dielectric loss and ac-conductivity enhancement in p-type Ag-doped SnO2 compounds. J. Alloys Compd. 2021, 870, 159515. [Google Scholar] [CrossRef]
  17. Narzary, R.; Dey, B.; Chouhan, L.; Kumar, S.; Ravi, S.; Srivastava, S.K. Optical band gap tuning, zero dielectric loss and room temperature ferromagnetism in (Ag/Mg) co-doped SnO2 compounds for spintronics applications. Mater. Sci. Semicond. Process. 2022, 147, 106477. [Google Scholar] [CrossRef]
  18. Beltrán, J.I.; Monty, C.; Balcells, L.; Martínez-Boubeta, C. Possible d0 ferromagnetism in MgO. Solid State Commun. 2009, 149, 1654–1657. [Google Scholar] [CrossRef]
  19. Graf, T.; Goennenwein, S.T.B.; Brandt, M.S. Prospects for carrier-mediated ferromagnetism in GaN. Phys. Status Solidi (B) 2003, 239, 277–290. [Google Scholar] [CrossRef]
  20. Morozov, I.G.; Belousova, O.V.; Belyakov, O.A.; Parkin, I.P.; Sathasivam, S.; Kuznetcov, M.V. Titanium nitride room-temperature ferromagnetic nanoparticles. J. Alloys Compd. 2016, 276, 266–276. [Google Scholar] [CrossRef]
  21. Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-like two-dimensional materials. Chem. Rev. 2013, 113, 3766–3798. [Google Scholar] [CrossRef] [PubMed]
  22. Si, C.; Zhou, J.; Sun, Z. Half-metallic ferromagnetism and surface functionalization-induced metal–insulator transition in graphene-like two-dimensional Cr2C crystals. ACS Appl. Mater. Interfaces 2015, 7, 17510–17515. [Google Scholar] [CrossRef] [PubMed]
  23. Feng, P.; Zhang, S.; Liu, D.; Gao, M.; Ma, F.; Yan, X.-W.; Xie, Z.Y. Achieving high-temperature ferromagnetism by means of magnetic ion dimerization in the graphene-like Mn2N6C6 monolayer. J. Phys. Chem. C 2022, 126, 10139–10144. [Google Scholar] [CrossRef]
  24. Zhang, H.; Liao, Z.; Xia, B.; Herng, T.S.; Ding, J.; Gao, D. Ferromagnetism of Mn-N4 architecture embedded graphene. J. Phys. D Appl. Phys. 2022, 55, 225001. [Google Scholar] [CrossRef]
  25. Wen, J.Q.; Tong, X.; Lei, Y.T.; Tian, P.H.; Wu, H.; He, W.L. Ferromagnetism induced by C doped graphene-like ZnO films with different concentrations. Solid State Commun. 2019, 299, 11366. [Google Scholar] [CrossRef]
  26. Wen, J.; Lin, P.; Han, Y.; Li, N.; Chen, G.; Bai, L.; Guo, S.; Wu, H.; He, W.; Zhang, J. Insights into enhanced ferromagnetic activity of P doping graphene-ZnO monolayer with point defects. Mater. Chem. Phys. 2021, 270, 124855. [Google Scholar] [CrossRef]
  27. Bafekry, A. Graphene-like BC6N single-layer: Tunable electronic and magnetic properties via thickness, gating, topological defects, and adatom/molecule. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 118, 113850. [Google Scholar] [CrossRef]
  28. Han, K.H.; Spemann, D.; Esquinazi, P.; Hohne, R.; Riede, V.; Butz, T. Ferromagnetic spots in graphite produced by proton irradiation. Adv. Mater. 2003, 15, 1719–1722. [Google Scholar] [CrossRef]
  29. Esquinazi, P.; Spemann, D.; Hohne, R.; Setzer, A.; Han, K.H.; Butz, T. Induced magnetic ordering by proton irradiation in graphite. Phys. Rev. Lett. 2003, 91, 227201. [Google Scholar] [CrossRef] [Green Version]
  30. Rode, A.V.; Gamaly, A.G.; Christy, J.G.; Fitz Gerald, S.T.; Hyde, R.G.; Elliman, B.; Luther-Davies, A.I.; Veinger, E.G.; Androulakis, J.; Giapintzakis, J. Unconventional magnetism in all-carbon nanoform. Phys. Rev. B Condens. Matter Mater. Phys. 2004, 70, 054407. [Google Scholar] [CrossRef]
  31. Castro, E.V.; Peres, N.M.R.; Stauber, T.; Silva, N.A.P. Low-density ferromagnetism in biased bilayer graphene. Phys. Rev. Lett. 2008, 100, 186803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Antonov, V.E.; Bashkin, I.O.; Khasanov, S.S.; Moravsky, A.P.; Morozov, Y.G.; Shulga, Y.M.; Ossipiyan, Y.A.; Ponyatovsky, E.G. Magnetic ordering in hydrofullerene C60H24. J. Alloys Compd. 2002, 330–332, 365–368. [Google Scholar] [CrossRef]
  33. Zhou, J.; Wang, Q.; Sun, Q.; Chen, X.S.; Kawazoe, Y.; Jena, P. Ferromagnetism in semihydrogenated graphene sheet. Nano Lett. 2009, 9, 3867–3870. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, Y.; Huang, Y.; Song, Y.; Zhang, X.; Ma, Y.; Liang, J. Room-temperature ferromagnetism of graphene. Nano Lett. 2009, 9, 220–224. [Google Scholar] [CrossRef] [PubMed]
  35. Qin, S.; Guo, X.; Cao, Y.; Ni, Z.; Xu, Q. Strong ferromagnetism of reduced graphene oxide. Carbon 2014, 78, 559–565. [Google Scholar] [CrossRef]
  36. Sinha, A.; Ali, A.; Thakur, A.D. Ferromagnetism in graphene oxide. Mater. Today Proc. 2021, 46, 6230–6233. [Google Scholar] [CrossRef]
  37. Błoński, P.; Tucek, J.; Sofer, Z.; Mazanek, V.; Petr, M.; Pumera, M.; Otyepka, M.; Zbořil, R. Doping with graphitic nitrogen triggers ferromagnetism in graphene. J. Am. Chem. Soc. 2017, 139, 3171–3180. [Google Scholar] [CrossRef] [Green Version]
  38. Ning, G.; Xu, C.; Hao, L.; Kazakova, O.; Fan, Z.; Wang, H. Ferromagnetism in nanomesh graphene. Carbon 2013, 51, 390–396. [Google Scholar] [CrossRef]
  39. Sarkar, S.K.; Raul, K.K.; Pradhan, S.S.; Basu, S.; Nayak, A. Magnetic properties of graphite oxide and reduced graphene oxide. Phys. E Low-Dimens. Syst. Nanostruct. 2014, 64, 78–82. [Google Scholar] [CrossRef]
  40. Liu, Y.; Tang, N.; Wan, X.; Feng, Q.; Li, M.; Xu, Q.; Liu, F.; Du, Y. Realization of ferromagnetic graphene oxide with high magnetization by doping graphene oxide with nitrogen. Sci. Rep. 2013, 3, 2566. [Google Scholar] [CrossRef]
  41. Qin, S.; Xu, Q. Room temperature ferromagnetism in N2 plasma treated graphene oxide. J. Alloys Compd. 2017, 692, 332–338. [Google Scholar] [CrossRef]
  42. Khurana, G.; Kumar, N.; Kotnala, R.K.; Nautiyal, T.; Katiyar, R.S. Temperature tuned defect induced magnetism in reduced graphene oxide. Nanoscale 2013, 5, 3346–3351. [Google Scholar] [CrossRef] [PubMed]
  43. Raj, K.G.; Joy, P.A. Ferromagnetism at room temperature in activated graphene oxide. Chem. Phys. Lett. 2014, 605–606, 89–92. [Google Scholar] [CrossRef]
  44. Rao, C.N.R.; Ramakrishna Matte, H.S.S.; Subrahmanyam, K.S. Synthesis and selected properties of graphene and graphene mimics. Acc. Chem. Res. 2013, 46, 149–159. [Google Scholar] [CrossRef] [PubMed]
  45. Bagani, K.; Ray, M.K.; Satpati, B.; Ray, N.R.; Sardar, M.; Banerjee, S. Contrasting magnetic properties of thermally and chemically reduced graphene oxide. J. Phys. Chem. C 2014, 118, 13254–13259. [Google Scholar] [CrossRef]
  46. Lee, D.; Seo, J.; Zhu, X.; Cole, J.M.; Su, H. Magnetism in graphene oxide induced by epoxy groups. Appl. Phys. Lett. 2015, 106, 172402. [Google Scholar] [CrossRef]
  47. Kim, S.W.; Kim, H.K.; Lee, K.; Roh, K.C.; Han, J.T.; Kim, K.B.; Lee, S.; Jung, M.-H. Studying reduction of graphene oxide with magnetic measurements. Carbon 2019, 142, 373–378. [Google Scholar] [CrossRef]
  48. Fu, L.; Wang, Y.; Zhang, K.; Zhang, W.; Chen, J.; Deng, Y.; Du, Y.; Tang, N. Realization of ambient-stable room-temperature ferromagnetism by low-temperature annealing of graphene oxide nanoribbons. ACS Nano 2019, 13, 6341–6347. [Google Scholar] [CrossRef]
  49. Nair, R.R.; Sepioni, M.; Tsai, I.-L.; Lehtinen, O.; Keinonen, J.; Krasheninnikov, A.V.; Thomson, T.; Geim, A.K.; Grigorieva, I.V. Spin-half paramagnetism in graphene induced by point defects. Nat. Phys. 2012, 8, 199–202. [Google Scholar] [CrossRef] [Green Version]
  50. Zhang, X.; Li, G.; Li, Q.; Shaikh, M.S.; Li, Z. The pure paramagnetism in graphene oxide. Results Phys. 2021, 26, 104407. [Google Scholar] [CrossRef]
  51. Vasiliev, V.P.; Manzhos, R.A.; Krivenko, A.G.; Kabachkov, E.N.; Shulga, Y.M. Nitrogen-enriched carbon powder prepared by ball-milling of graphene oxide with melamine: An efficient electrocatalyst for oxygen reduction reaction. Mendeleev Commun. 2021, 31, 529–531. [Google Scholar] [CrossRef]
  52. Vasiliev, V.P.; Manzhos, R.A.; Kochergin, V.K.; Krivenko, A.G.; Kabachkova, E.N.; Kulikov, A.V.; Shulga, Y.M.; Gutsev, G.L. A Facile synthesis of noble-metal-free catalyst based on nitrogen doped graphene oxide for oxygen reduction reaction. Materials 2022, 15, 821. [Google Scholar] [CrossRef] [PubMed]
  53. Shulga, Y.M.; Baskakov, S.A.; Smirnov, V.A.; Shulga, N.Y.; Belay, K.G.; Gutsev, G.L. Graphene oxide films as separators of polyaniline-based supercapacitors. J. Power Sources 2014, 245, 33–36. [Google Scholar] [CrossRef]
  54. Stylianakis, M.M.; Spyropoulos, G.D.; Stratakis, E.; Kymakis, E. Solution-processable graphene linked to 3,5-dinitrobenzoyl as an electron acceptor in organic bulk heterojunction photovoltaic devices. Carbon 2012, 50, 5554–5561. [Google Scholar] [CrossRef]
  55. Shahriary, L.; Athawale, A. Graphene oxide synthesized by using modified hummers approach. Int. J. Renew. Energy Environ. Eng. 2014, 2, 58–63. [Google Scholar]
  56. Saleem, H.; Haneef, M.; Abbasi, H.Y. Synthesis route of reduced graphene oxide via thermal reduction of chemically exfoliated graphene oxide. Mater. Chem. Phys. 2018, 204, 1–7. [Google Scholar] [CrossRef]
  57. Ikram, R.; Jan, B.M.; Ahmad, W. An overview of industrial scalable production of graphene oxide and analytical approaches for synthesis and characterization. J. Mater. Res. Technol. 2020, 9, 11587–11610. [Google Scholar] [CrossRef]
  58. Meier, R.J.; Maple, J.R.; Hwang, M.-J.; Hagler, A.T. Molecular modeling urea- and melamine-formaldehyde resins. 1. A force field for urea and melamine. J. Phys. Chem. 1995, 99, 5445–5456. [Google Scholar] [CrossRef]
  59. Kitson, R.E.; Griffith, N.E. Infrared absorption band due to nitrile stretching vibration. Anal. Chem. 1952, 24, 334–337. [Google Scholar] [CrossRef]
  60. Dows, D.A.; Haim, A.; Wilmarth, W.K. Infra-red spectroscopic detection of bridging cyanide groups. J. Inorg. Nucl. Chem. 1961, 21, 33–37. [Google Scholar] [CrossRef]
  61. Kumar, M.; Chung, J.S.; Kong, B.-S.; Kim, E.J.; Hur, S.H. Synthesis of graphene–polyurethane nanocomposite using highly functionalized graphene oxide as pseudo-crosslinker. Mater. Lett. 2013, 106, 319–321. [Google Scholar] [CrossRef]
  62. Pimentel, G.C.; McClellan, A.L. Hydrogen bonding. Annu. Rev. Phys. Chem. 1971, 22, 347–385. [Google Scholar] [CrossRef]
  63. Zhang, C.; Dabbs, D.M.; Liu, L.-M.; Aksay, I.A.; Car, R.; Selloni, A. Combined effects of functional groups, lattice defects, and edges in the infrared spectra of graphene oxide. J. Phys. Chem. C 2015, 119, 18167–18176. [Google Scholar] [CrossRef]
  64. Lazar, P.; Mach, R.; Otyepka, M. Spectroscopic fingerprints of graphitic, pyrrolic, pyridinic, and chemisorbed nitrogen in N-doped graphene. J. Phys. Chem. C 2019, 123, 10695–10702. [Google Scholar] [CrossRef]
  65. Lesiak, B.; Kövér, L.; Tóth, J.; Zemek, J.; Jiricek, P.; Kromka, A.; Rangam, N. C sp2/sp3 hybridisations in carbon nanomaterials—XPS and (X) AES study. Appl. Surf. Sci. 2018, 452, 223–231. [Google Scholar] [CrossRef]
  66. Shulga, Y.M.; Kabachkov, E.N.; Korepanov, V.I.; Khodos, I.I.; Kovalev, D.Y.; Melezhik, A.V.; Tkachev, A.G.; Gutsev, G.L. Concentration of C sp3 atoms and other properties of an activated carbon with over 3000 m2/g BET surface area. Nanomaterials 2021, 11, 1324. [Google Scholar] [CrossRef]
  67. Wang, B.; Fielding, A.J.; Dryfe, R.A.W. Electron paramagnetic resonance as a structural tool to study graphene oxide: Potential dependence of the EPR response. J. Phys. Chem. C 2019, 123, 22556–22563. [Google Scholar] [CrossRef]
  68. Zener, C. Interaction between the d-shells in the transition metals. II. Ferromagnetic compounds of manganese with perovskite structure. Phys. Rev. 1951, 82, 403–405. [Google Scholar] [CrossRef]
Figure 1. SEM micrographs of GO, and NDCNM.
Figure 1. SEM micrographs of GO, and NDCNM.
Molecules 27 07698 g001
Figure 2. IR spectra of graphene oxide (1), NDCNM (2) and melamine (3).
Figure 2. IR spectra of graphene oxide (1), NDCNM (2) and melamine (3).
Molecules 27 07698 g002
Figure 3. Survey XPS spectra of initial GO (1), sample of GO ground in a planetary mill (2) and NDCNM powder (3).
Figure 3. Survey XPS spectra of initial GO (1), sample of GO ground in a planetary mill (2) and NDCNM powder (3).
Molecules 27 07698 g003
Figure 4. N1s (a) and C1s (b) high resolution spectra of NDCNM.
Figure 4. N1s (a) and C1s (b) high resolution spectra of NDCNM.
Molecules 27 07698 g004
Figure 5. (a)—ESR spectra of initial GO (1) and ground GO (2); (b)—ESR spectra of NDCNM powder (3) (top, in a wide range of magnetic field; bottom, in a narrow range of magnetic field). Spectra were obtained at a room temperature.
Figure 5. (a)—ESR spectra of initial GO (1) and ground GO (2); (b)—ESR spectra of NDCNM powder (3) (top, in a wide range of magnetic field; bottom, in a narrow range of magnetic field). Spectra were obtained at a room temperature.
Molecules 27 07698 g005
Figure 6. Specific magnetization σ (emu/g) as a function of the magnetic field H (Oe).
Figure 6. Specific magnetization σ (emu/g) as a function of the magnetic field H (Oe).
Molecules 27 07698 g006
Table 1. Elemental composition of the studied samples (wt.%).
Table 1. Elemental composition of the studied samples (wt.%).
SampleCHNS
Melamine *28.574.7666.670.000
melamine (grinding)28.714.54465.240.042
graphene oxide52.132.1131.070.070
GO (grinding)53.282.2771.130.065
NDCNM52.472.4736.580.033
*—calculated according to the chemical formula of melamine (C3N6H6).
Table 2. Elemental composition (at.%) of the samples as determined from XPS spectra.
Table 2. Elemental composition (at.%) of the samples as determined from XPS spectra.
SampleElement
CNOS
graphene oxide74.30.323.51.9
GO (grinding)76.80.023.00.2
NDCNM76.75.517.40.4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vasiliev, V.P.; Kabachkov, E.N.; Kulikov, A.V.; Manzhos, R.A.; Morozov, I.G.; Shulga, Y.M. Unexpected Room Temperature Ferromagnetism of a Ball-Milled Graphene Oxide—Melamine Mixture. Molecules 2022, 27, 7698. https://doi.org/10.3390/molecules27227698

AMA Style

Vasiliev VP, Kabachkov EN, Kulikov AV, Manzhos RA, Morozov IG, Shulga YM. Unexpected Room Temperature Ferromagnetism of a Ball-Milled Graphene Oxide—Melamine Mixture. Molecules. 2022; 27(22):7698. https://doi.org/10.3390/molecules27227698

Chicago/Turabian Style

Vasiliev, Vladimir P., Eugene N. Kabachkov, Alexander V. Kulikov, Roman A. Manzhos, Iurii G. Morozov, and Yury M. Shulga. 2022. "Unexpected Room Temperature Ferromagnetism of a Ball-Milled Graphene Oxide—Melamine Mixture" Molecules 27, no. 22: 7698. https://doi.org/10.3390/molecules27227698

Article Metrics

Back to TopTop