Next Article in Journal
Observation of Josephson-like Tunneling Junction Characteristics and Positive Magnetoresistance in Oxygen Deficient Nickelate Films of Nd0.8Sr0.2NiO3−δ
Next Article in Special Issue
Growing Tungsten Nanophases on Carbon Spheres Doped with Nitrogen. Behaviour as Electro-Catalysts for Oxygen Reduction Reaction
Previous Article in Journal
Effect of a Nanostructured Titanium Surface on Gingival Cell Adhesion, Viability and Properties against P. gingivalis
Previous Article in Special Issue
Modified Mesoporous Carbon Material (Pb-N-CMK-3) Obtained by a Hard-Templating Route, Dicyandiamide Impregnation and Electrochemical Lead Particles Deposition as an Electrode Material for the U(VI) Ultratrace Determination
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Porous Expanded Graphite: Thermal Shock vs. Programmable Heating

by
Alexander G. Bannov
1,*,
Arina V. Ukhina
2,
Evgenii A. Maksimovskii
3,
Igor Yu. Prosanov
2,
Artyom A. Shestakov
1,
Nikita I. Lapekin
1,
Nikita S. Lazarenko
1,
Pavel B. Kurmashov
1 and
Maksim V. Popov
1,4
1
Department of Chemistry and Chemical Engineering, Faculty of Mechanical Engineering, Novosibirsk State Technical University, 630092 Novosibirsk, Russia
2
Institute of Chemistry of Solid State and Mechanochemistry, Siberian Branch, Russian Academy of Sciences, 630092 Novosibirsk, Russia
3
Institute of Inorganic Chemistry, Siberian Branch, Russian Academy of Sciences, 630090 Novosibirsk, Russia
4
N.D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Materials 2021, 14(24), 7687; https://doi.org/10.3390/ma14247687
Submission received: 3 November 2021 / Revised: 8 December 2021 / Accepted: 10 December 2021 / Published: 13 December 2021
(This article belongs to the Special Issue Development and Applications of Advanced Carbon Materials)

Abstract

:
Highly porous expanded graphite was synthesized by the programmable heating technique using heating with a constant rate (20 °C/min) from room temperature to 400–700 °C. The samples obtained were analyzed by scanning electron microscopy, energy-dispersive spectroscopy, low-temperature nitrogen adsorption, X-ray photoelectron spectroscopy, Raman spectroscopy, thermogravimetry, and differential scanning calorimetry. A comparison between programmable heating and thermal shock as methods of producing expanded graphite showed efficiency of the first one at a temperature 400 °C, and the surface area reached 699 and 184 m2/g, respectively. The proposed technique made it possible to obtain a relatively higher yield of expanded graphite (78–90%) from intercalated graphite. The experiments showed the advantages of programmable heating in terms of its flexibility and the possibility to manage the textural properties, yield, disorder degree, and bulk density of expanded graphite.

Graphical Abstract

1. Introduction

Expanded graphite (EG) is used for the production of sealing materials [1], graphite paper [2,3,4], protection coatings [5], fillers for polymer composites [6,7,8], adsorbent [9], electrode material for supercapacitors [10,11], cathodes for aluminum-ion batteries [12], components of antibacterial materials [13,14], adsorbents [15], catalysts [16], etc.
There are two main methods of manufacturing EG, thermal expansion [17,18] and liquid-phase expansion [19]. Thermal expansion is usually used to produce EG from intercalated graphite [20]. Rapid heating leads to the decomposition of intercalated graphite compounds and induces strong formation of a gas phase, which produces a porous cellular structure [21]. The relatively high surface area of the EG determines its applications in adsorption, etc. [22]. In [23], liquid phase expansion was carried out using the mixture of H2SO4 and H2O2 (9:1 vol. ratio) and flake graphite.
The porosity of expanded graphite strongly depends on the method of intercalation. For example, van Heerden et al. [24] carried out a comparison of three intercalation methods (FeCl3 gas phase intercalation, electrochemical intercalation, and Hummers method) with regard to the characteristics of EG. Gas phase intercalation with FeCl3 was carried out using the mixture of graphite and anhydrous FeCl3 (1:2 wt. ratio), which were dried (50 °C, 2 h) and heated (300 °C, 25 h). It was found that electrochemical intercalation in sulfuric acid electrolyte and the Hummers method led to the formation of indistinguishable products with similar surface areas of 18.7 and 21.7 m2/g, respectively. In [25], the authors carried out heating of graphite intercalation compounds (GICs) from 800 to 900 °C for 20 s, leading to exfoliation of graphite oxide and expanded graphite, but the change in specific surface area was relatively low. In [26,27], it was concluded that the surface area of graphite intercalated with sulfuric acid was low enough (45 m2/g). Goudarzi et al. [28] noted that the particle size of GICs and temperature of exfoliation significantly affected the pore size of expanded graphite. Textural characteristics increased when the temperature increased from 700 to 900 °C. The increase in GIC particle size induces a drop in specific pore area of EG.
Usually, the surface area of expanded graphite is relatively low; therefore, novel methods for its synthesis must be developed in order to improve the important characteristics for further application. In this study, programmable heating was used to synthesize EG. The effects of various final temperatures used for heating intercalated graphite on the disorder degree, porosity, and chemical composition of EG were studied in detail. A comparison between thermal shock and programmable heating used for obtaining EG was done.

2. Experimental

2.1. Preparation of Expanded Graphite

Two sets of highly porous expanded graphite samples were synthesized.
The first set contained 4 samples, which were synthesized from 1 g of intercalated graphite (EG-350-50, Chemical Systems Ltd., Moscow, Russia). To synthesize the thermally expanded graphite and control its porosity, programmable heating, as the first synthesis technique, was used. In addition, a comparison of programmable heating with thermal shock, the second synthesis technique, was done.
The description of samples synthesized under different conditions is shown in Table 1.
Programmable heating was performed by using a certain temperature program set in a muffle furnace (SNOL, Utena, Lithuania) in which the Al2O3 crucible with intercalated graphite was placed. The sample was heated from room temperature (25 ± 1 °C) to maximum temperatures of 400, 500, 600, and 700 °C in air atmosphere. The heating rate was 20 °C/min, set by the controller of the furnace. The cooled samples were taken from the furnace and were analyzed.
The second method was thermal shock, which was used for comparison with programmable heating. The crucible with the sample was quickly placed into the hot furnace. Usually to synthesize EG, temperatures of 800–1200 °C are used [29,30,31]; therefore, to carry out the comparison, 2 temperatures were used, one lower than this range (400 °C) and one within this range (1000 °C), taking into account that the temperature gradients that influence EG are around 375 and 975 °C/s. Each sample was kept in the furnace for 10 min at the given temperature, then the sample was taken out and cooled down in air atmosphere without additional intensification of cooling.
The appearance of samples obtained by programmable heating and thermal shock are shown in Figure 1.

2.2. Characterization

The EG obtained was characterized by the following method. The bulk density of the sample was measured according to GOST 25699.14-93 by compaction of the bed of EG placed into a 10 mL measuring cylinder under the action of vibration until the constant height of the bed was reached.
The yield of EG (%) synthesized using the 2 techniques was determined by weighing the samples using an analytical balance (Sartogosm, Sankt-Peterburg, Russia) as the ratio of sample weight after synthesis to sample weight before synthesis.
The morphology of the surface of samples obtained was investigated by a S-3400N scanning electron microscope (SEM) (Hitachi, Japan). The add-on for energy-dispersive X-ray spectroscopy (EDX) (Oxford Instruments, Abingdon, UK) was used to analyze elements present in the samples.
X-ray diffraction was performed on a DRON-3 diffractometer (Polysorb, Russia) using CuKα radiation (λ = 1.54 Å). The number of carbon layers (n) in a crystallite along the c-axis direction was calculated by Equation (1) [32,33]:
n = L c d 002 + 1
where Lc is the crystallite size and d002 is the interlayer distance. Lc parameter was calculated using Sherrer’s formula with values of K = 0.9 [34].
The graphitization degree (g, %) was calculated by the Maire–Mering Equation (2) [35].
g = 3.44 d 002 3.44 3.354
Raman spectroscopy was performed using T64000 Horiba Jobin Yvon (Japan) (λ = 514 nm) spectrometer.
The study of the chemical composition of sample surfaces was done using X-ray photoelectron spectroscopy (XPS). The analysis was performed using a surface nano analysis spectrometer (SPECS GmbH, Berlin, Germany) equipped with a PHOIBOS-150 hemispherical analyzer and an XR-50M X-ray radiation source with double Al/Ag anode. Al Kα monochromized radiation (h* = 1486.74 eV) was used for the study. The relative concentrations of elements in zones of analysis were determined based on the integral intensities of XPS spectral lines, taking into account the section of photoionization of corresponding terms. The spectra were deconvoluted by individual components. After subtracting the baseline by the Shirley method, the experimental curve was decomposed to a set of lines corresponding to the photoemission of electrons from atoms in various surrounding chemicals. CasaXPS software (2.3.24, Teignmouth, U.K.) was used for data treatment. The shape of peaks was approximated with a symmetric function obtained by multiplying the Lorenz and Gauss functions.
The texture characteristics of the obtained carbon materials were studied using low-temperature nitrogen adsorption at 77 K using Quantachrome NOVA 1000 e adsorption unit (Quantachrome Instruments, Boynton Beach, Florida, USA). Specific surface area was calculated using the BET method. The surface area of the pores remaining after filling the micropores with the adsorbate was calculated by the comparative t-method (de Boor and Lippens). Pore size distribution was calculated using the Barrett–Joyner–Halenda (BJH) method. The samples were preliminarily subjected to degassing in a vacuum at 50 Torr for 5 h at 300 °C. The mass of one sample was around 0.01 g.
Thermogravimetry (TG) and differential scanning calorimetry (DSC) were carried out using a STA 449C thermal analyzer (Netzsch, Selb, Germany). The heating was conducted in an Al2O3 crucible with a tip. The temperature of exfoliation of intercalated graphite was determined in pure argon. The stability of EG samples toward oxidation was analyzed at a heating rate of 10 °C/min in a mixture of argon and oxygen.

3. Results and Discussion

Figure 2 shows the TG and DSC curves of intercalated graphite used for preparing EG.
The mass of the sample decreased during heating and there was a sharp drop in the TG curve at 193 °C. According to the DSC curve, the onset of the exothermic peak attributed to exfoliation of the intercalated graphite (as a result of decomposition of graphite intercalation compounds) was 267 °C, and this temperature can be treated as the beginning of thermal exfoliation [36]. The maximum of DSC peak was detected at 360 °C. It was not possible to determine the end of the DSC peak, since the curve was noisy, and this effect was repeated many times with intercalated graphite. It is interesting that the mass loss continued during heating above 400 °C, although the value of loss within a temperature range 400–700 °C was 2.87%. This indicates the occurrence of additional processes in graphite after its exfoliation.
SEM images of the samples subjected to programmable heating and thermal shock are shown in Figure 3 with SEM images of initial intercalated graphite (Figure 3a,b) shown for comparison. Intercalated graphite showed morphology typical for graphite. Heating of intercalated graphite induced the deformation of porous particles (Figure 3c,d) of the sample, with the appearance of cracks on the surface (Figure 3e,f).
Table 2 shows the EDX data of intercalated graphite and TEG1-400. The initial intercalated graphite contained the following elements: mainly C (87.31 at.%), O (11.06 at.%), and S (1.14 at.%), and traces of Al (0.06 at.%), Si (0.18 at.%), Cr (0.05 at.%), Fe (0.07 at.%), and Cu (0.12 at.%).
The presence of sulfur and oxygen is mainly attributed to graphite intercalated compounds, the content of which decreased after thermal expansion of the sample. The composition of intercalated graphite can indicate that the sample is represented by graphite bisulfate. The presence of chromium in the sample makes it possible to suggest that K2Cr2O7 was used as an oxidizer for synthesis. According to data on thermal analysis, the DSC peak attributed to the decomposition of GICs (graphite bisulfate) is exothermic, whereas some studies [37,38] reported that it had an endothermic nature. An exothermic DSC peak appearing as a result of graphite bisulfate decomposition was also found in [39]. Taking into account that the sample expansion was strong enough, there was also small amount of residual elements present in the TEG1-400, attributed to the GICs (S, 0.29 wt.%; O, 5.60 wt.%). The presence of silicon, iron, and copper can be explained by their trace amounts in graphite used for commercial intercalated graphite. The sulfur content in IG is in agreement with EDX data reported in [29].
As can be seen, the EG sample had a porous structure that was more well-developed compared to the intercalated graphite. Figure 4 shows the dependence of the bulk density of EG synthesized by programmable heating on the synthesis temperature. The bulk density of the samples obtained by programmable heating ranged from 0.017 to 0.058 g/cm3. Increasing the temperature of synthesis induced a decrease in bulk density; the highest bulk density value (0.058 g/cm3) was achieved at 400 °C. In addition, increasing the temperature above 500 °C led to an almost constant bulk density value (0.017–0.018 g/cm3) independent of the temperature. Temperatures above 700 °C were not used for the programmable heating method because of strong oxidation of EG beginning from 650–700 °C. The thermal shock method showed similar bulk density values (TEG2-400, 0.016 g/cm3; TEG2-1000, 0.019 g/cm3), which are lower than that of TEG1-400 (0.058 g/cm3) due to higher pressure acting on the structural defects of the material during expansion at relatively high heating rates of ~400 and 1000 °C/s.
The bulk density values were close to the values obtained in [28] for expanded graphite synthesized from GICs with 200 mesh. This mesh size is close enough to the particle size of IG, which was in a range of 100–200 µm.
The data on bulk density, yield, and texture characteristics are shown in Table 3.
One of the important characteristics showing the efficiency of the synthesis process is the yield of EG. Usually, the most of authors do not pay attention to yield, however it is extremely important for chemical engineering and further scale up of technology. The yield was in the range of 77.8–89.2% and decreased with increasing temperature. The highest yield was reached at 400 °C, whereas the bulk density was relatively high, which is inappropriate for some applications (adsorption [40], catalysis [41,42], sealing materials [1], etc.) where this value must be lower. The bulk density of TEG1-500 sample is low enough, but the surface area and yield were lower compared to sample prepared at 400 °C. The difference in yield TEG1-400 and TEG1-500 was 10.2% and it is high enough. Therefore TEG1-400 sample is better compared to other samples from technological and economical points of view.
Specific surface area is linked to bulk density and yield. In particular, the highest surface area of 699 m2/g corresponded to TEG1-400, and it decreased with increasing temperature. The bulk density of the samples synthesized at 500–700 °C by programmable heating was almost the same, but the specific surface area changed significantly. The samples obtained using programmable heating and thermal shock were predominantly mesoporous, with a fraction of micropores within 26–41%, and for TEG1-400, TEG1-500, TEG1-600, and TEG1-700, it was 37.3, 26.0, 33.1, and 41.4%, respectively. Thus, increasing temperature induces a decreasing fraction of mesopores. Probably increasing the temperature above 500 °C leads to stronger “opening” of mesopores and micropores, leading to decreased fractions and reduced total surface area. The fraction of both micropores and mesopores in programmable heating decreases with increasing temperature. For thermal shock, the dependence is opposite, i.e., both micropores and mesopores increase.
The differences in porous structure of the samples can be seen in SEM images shown in Figure 3 and Figure 5. The samples obtained using thermal shock look more porous compared to those using programmable heating. However, the data on surface area showed that the difference between the porous structure of samples is significant.
Comparing the textural characteristics of samples obtained by programmable heating and thermal shock, we can see that the samples synthesized by the former had a higher specific surface area than the latter at higher yield. The yield of samples TEG2-400 and TEG2-1000 was 77.4 and 76.9%, respectively, due to the higher driving force of the thermal expansion process at high temperature gradients.
The XRD pattern of the initial commercial intercalated graphite is shown in Figure 6a.
The spectrum is represented by a wide peak around 2θ = 26°, which consists of two main peaks: 2θ = 26.5°, corresponding to graphite (d002 = 3.368 Å), and 2θ = 26.0°, with an interlayer distance lower than that in pure graphite, corresponding to graphite bisulfate [43]. The EDX and XRD data suggest that the sample of initial intercalated graphite is graphite bisulfate obtained by using K2Cr2O7 as an oxidizing agent. This is also confirmed by XRD data presented in [43], showing a shift of the 002 peak to below 2θ = 26° with an asymmetrical peak appearing in the spectrum. The XRD spectrum of TEG1-400 is represented only by graphite phase and a narrow peak at 2θ = 26.5° (Figure 6b).
There is a link between the number of carbon layers in the crystallite along the c direction, the crystallite size Lc, and the specific surface area (Figure 7a). The curves of pore size distribution are presented in Figure 7b–g.
This shows that the specific area decreases with the growth of crystallite size, and the value of Lc = 290 Å corresponds to the highest surface area (699 m2/g). The data of average pore size confirm that the pore diameter of EG decreases when increasing temperature using programmable heating. Thermal shock led to the increase in average pore diameter when increasing temperature.
Isotherms of adsorption are presented in Supplementary Materials (Figure S2). Based on the type of adsorption and desorption isotherms, it can be concluded that all EG samples have an inhomogeneous porous structure. All EG samples are characterized by a steep rise of the isotherm in the region P/P0 < 0.4 and up to P/P0 = 0.99, which indicates the presence of micropores in the structure, which can have a positive effect on the sorption properties of the material. An insignificant loop of the capillary condensation hysteresis of the H4 type is observed when increasing the relative pressure in the region P/P0 = 0.4–0.99 due to the presence of slit-shaped mesopores [44]. The isotherms of the analyzed EG samples are difficult to classify according to the generally accepted classification of physical sorption isotherms adopted by IUPAC. The type of analyzed isotherms is most closely matched with IVa type from the proposed updated qualification of isotherms [45]. Type IV is typical for mesoporous adsorbents (e.g., mesoporous molecular sieves). In the case of type IVa isotherms, capillary condensation is accompanied by a slight hysteresis of the H4 type. This occurs when the pore width exceeds a certain critical width, which depends on the adsorption system and temperature, which is typical for cylindrical, narrowed, and slit-like pores with a diameter of ~4 nm.
The FTIR spectra of typical TEG1-400 and IG samples are shown in Figure 8.
The IG and expanded graphite spectra show the presence of O–H vibrations (3500 cm−1), C=C stretching vibrations (1642 cm−1), and CO stretching vibrations (1750 cm−1) [29,46]. The IG spectrum shows an C=S stretching band at 1050 cm−1, the intensity of which decreased after exfoliation [47].
Raman spectra and the parameters of spectra of EG are presented in Figure 9 and Table 4. Raman spectrum of IG is presented in Supplementary Materials (Figure S1).
There were D and G bands typical for graphite materials presented in Raman spectra corresponded to disorder in graphite and E2g mode of graphite related to sp2 bonded atoms [48]. The position of G peak remained unchanged for the entire set of samples. D and G peaks became narrow when increasing the temperature of graphite expansion. The ratio of intensity of peaks I (D)/I (G) was low enough indicating perfect crystallinity of expanded graphites. The highest I (D)/I (G) ratio attributed to defectiveness of EG was shown by TEG1-400 and it decreased as the temperature rise. The difference of I (D)/I (G) ratio of samples TEG1-400 and TEG2-400 obtained using programmable heating and thermal shock is high enough (0.1 and 0.014, respectively), and it is in agreement with textural characteristics shown in Table 3. The data corresponded to samples obtained by thermal shock showed the deeper removal of GICs from the material making it closer to pure graphite, i.e., material with high graphitization degree. The difference of I (D)/I (G) ratio of TEG1-400 and TEG2-400 (0.1 and 0.014, respectively) samples is in agreement with data of textural characteristics. It was supposed that the higher expansion degree the higher defectiveness. The gas phase formed as a result of decomposition of GICs passes through the certain defect regions in the sample obtained using thermal shock. The number of regions through which the gas phase penetrates in the sample TEG1-400 is higher since its pressure is not so high as in thermal shock. It means that the samples obtained using programmable heating will have higher defectiveness compared to thermal shock because of larger number of defect regions available for gas phase pathways. Higher temperature gradients lead to contribution of only limited defects which were immediately expanded and the subsequent drop in driving force decelerates the expansion.
According to XPS, the typical C1s spectra of IG and EG consist of four main components corresponding to C=C (284.5 eV), C-O (286.5 eV), C=O (288 eV), and O-C=O (288.8 eV) groups (Figure 10) [49,50].
The fractions of these components for IG were 96.7%, 2.8%, 0.3%, and 0.2%, respectively; after exfoliation (TEG1-400), they were 92.9%, 5.6%, 1.1%, and 0.4%, respectively. Exfoliation led to an increase in the C=C component of the C1s spectrum attributed to carbon bonding with carbon in sp2 hybridization. It also induced the removal of C-O and C=O groups from the intercalated graphite during heating, whereas the concentration of carboxyls decreased slightly. This confirms the thermodynamic stability of carboxylic groups with respect to heating. However, some articles reported that the temperature range of the release of carboxylic groups was 100–250 °C [51].
The O1s spectrum of intercalated graphite is represented by a peak at 531.5 eV attributed to oxygen in SOx groups, a peak at 532.5 eV attributed to oxygen in C = O groups, and two peaks at 533.6 and 535.1–535.7 eV corresponding to OH groups and water, respectively (Figure 10c). This gives us additional confirmation of the presence of graphite bisulfate in the initial sample used for exfoliation.
The oxidation resistance of expanded graphite is an extremely important characteristic linked to defectiveness, specific surface area, and chemical composition. DSC makes it possible to estimate the dynamics of oxidation of EGs obtained using two methods under consideration. The results of the analysis of DSC data on EG thermal oxidation are shown in Table 5.
It can be seen that the oxidation process of EG has a dependence on the synthesis temperature for the samples obtained using programmable heating. The DSC curves of EG samples are shown in Figure 11.
For example, the TEG1-500 sample showed higher temperature oxidation at the beginning and the highest temperature at the maximum DSC peak. The increased synthesis temperature makes the DSC exothermic oxidation peak broader compared with the relatively narrow peaks for the TEG1-400 and TEG1-500 samples. The TEG1-400 sample showed the lowest heat release during oxidation of 9582 J/g. The EG obtained had a large surface area along with high graphitization degree, accompanied by high onset oxidation temperature, which makes it a good candidate for catalysis (catalyst support for high-temperature processes) and electrochemistry (as a material with good conductivity and porosity).
It is worth noting that the oxidation resistance of the EG obtained was high enough, higher than that of various porous graphites, and can be compared to the expanded graphite coated with SiC (expanded graphite +25 or 100 vol.% silane coupling agent) synthesized in [52], graphite foil obtained from expandable graphite modified with boric acid [53], and boron oxide modified graphite foil reported by Savchenko et al. [54].
According to the data presented above, programmable heating is an alternative way to improve the textural properties and to obtain porous expanded graphite. Table 6 shows a comparison of the specific surface area of expanded graphite obtained in this work with data reported in the literature. As can be seen, the EG synthesized in this study significantly exceeds other samples, and its specific surface area is about 100–200 m2/g higher than that of expanded graphite obtained from graphite oxide or graphene oxide, which makes this approach very attractive for practical applications. The high specific surface area of EG obtained by programmable heating can be explained by the relatively low heating rate, which makes it possible for the balloon walls in the cells to expand more gradually, while thermal shock is rapid heating that results in fast formation of expanded cells occurring simultaneously with the formation of defects in the graphite structure. These defects are paths for the gas phase formed as a result of the decomposition of GICs. Since thermal shock has an extremely high heating rate, these defective regions expand rapidly and the gas phase leaves the cell. This effect of “bleeding” of gas leads to the expansion of a certain small number of pores, leaving the remaining pores almost unexpanded. Compared with thermal shock, the heating rate used in programmable heating is low enough that it preserves some defects from rapid expansion and the gas phase stays in the balloon cell. Thus, the number of pores formed is higher and the specific surface area is high enough.
It is worth noting that the heating rate in the programmable heating method is an analogue of residence time, a value usually used in chemical engineering to describe the residence of reagent in a reactor or other apparatus. The higher the heating rate, the lower the residence time, and vice versa.
In conclusion, it can be noted that the method of programmable heating makes it possible to obtain highly porous material at relatively moderate temperatures (400–500 °C) with higher yield. The advantage of this method is the possibility to increase the time of exposure of intercalated graphite in the furnace. At the same time, this method is more flexible than thermal shock because it allows for the possibility to vary the textural properties, yield, and bulk density in a wide range.

4. Conclusions

In this study, the highly porous expanded graphite was synthesized by the programmable heating technique using heating with a constant rate (20 °C/min) from room temperature to 400–700 °C. A comparison between programmable heating and thermal shock as methods of producing expanded graphite showed its efficiency at an extremely low temperature (400 °C), and the surface area reached 699 and 184 m2/g. It was found that the novel method of production of expanded graphite from intercalated graphite made it possible to obtain a relatively higher yield (78–90%). The EG obtained using programmable heating had a large surface area along with high graphitization degree, accompanied by high onset oxidation temperature, which makes it a good candidate for catalysis and electrochemistry. The programmable heating led to obtaining predominantly mesoporous expanded graphites. The advantage of this method is the possibility to increase the time of exposure of intercalated graphite in the furnace. This method is more flexible than thermal shock because it allows for the possibility to vary the textural properties, yield, and bulk density in a wider range.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ma14247687/s1, Figure S1: Raman spectrum of the intercalated graphite, Figure S2: Isotherms of adsorption of expanded graphites: TEG1-400 (a), TEG1-500 (b), TEG1-600 (c), TEG1-700 (d), TEG2-400 (e), TEG2-1000 (f).

Author Contributions

Methodology, writing—original draft preparation, supervision, A.G.B., A.A.S., N.I.L., N.S.L.; writing—review and editing, A.G.B.; resources, M.V.P.; investigation, A.V.U., E.A.M., I.Y.P., P.B.K. All authors participated in the discussions of the results. All authors have read and agreed to the published version of the manuscript.

Funding

The work was funded by the State task of Ministry of Science and Higher Education of Russia (project no. FSUN-2020-0008).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sorokina, N.E.; Redchitz, A.V.; Ionov, S.G.; Avdeev, V.V. Different exfoliated graphite as a base of sealing materials. J. Phys. Chem. Solids 2006, 67, 1202–1204. [Google Scholar] [CrossRef]
  2. Pandey, I.; Tiwari, J.D.; Saha, T.K.; Khosla, A.; Furukawa, H.; Sekhar, P.K. Black carbon paper based polyanthraquinone coated exfoliated graphite for flexible paper battery. Microsyst. Technol. 2019, 3, 1–9. [Google Scholar] [CrossRef]
  3. Huang, L.; Rao, W.; Fan, L.; Xu, J.; Bai, Z.; Xu, W.; Bao, H. Paper electrodes coated with partially-exfoliated graphite and polypyrrole for high-performance flexible supercapacitors. Polymers 2018, 10, 135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Raghunandanan, A.; Yeddala, M.; Padikassu, P.; Pitchai, R. Partially Exfoliated Graphite Paper as Free-Standing Electrode for Supercapacitors. ChemistrySelect 2018, 3, 5032–5039. [Google Scholar] [CrossRef]
  5. Pajarito, B.B.; Cayabyab, C.A.L.; Costales, P.A.C.; Francisco, J.R. Exfoliated graphite/acrylic composite film as hydrophobic coating of 3D-printed polylactic acid surfaces. J. Coat. Technol. Res. 2019, 16, 1133–1140. [Google Scholar] [CrossRef]
  6. Berdyugina, I.S.; Steksova, Y.P.; Shibaev, A.A.; Maksimovskii, E.A.; Bannov, A.G. Thermal degradation of epoxy composites based on thermally expanded graphite and multiwalled carbon nanotubes. Russ. J. Appl. Chem. 2016, 89, 1447–1453. [Google Scholar] [CrossRef]
  7. Yim, Y.-J.; Park, S.-J. Effect of silver-plated expanded graphite addition on thermal and electrical conductivities of epoxy composites in the presence of graphite and copper. Compos. Part A Appl. Sci. Manuf. 2019, 123, 253–259. [Google Scholar] [CrossRef]
  8. Bannov, A.G.; Nazarenko, O.B.; Maksimovskii, E.A.; Popov, M.V.; Berdyugina, I.S. Thermal behavior and flammability of epoxy composites based on multi-walled carbon nanotubes and expanded graphite: A comparative study. Appl. Sci. 2020, 10, 6928. [Google Scholar] [CrossRef]
  9. He, J.; Song, L.; Yang, H.; Ren, X.; Xing, L. Preparation of Sulfur-Free Exfoliated Graphite by a Two-Step Intercalation Process and Its Application for Adsorption of Oils. J. Chem. 2017, 2017, 1–8. [Google Scholar] [CrossRef] [Green Version]
  10. Rachiy, B.I.; Budzulyak, I.M.; Ivanenko, E.A.; Revo, S.L. A composite of nanoporous carbon and thermally exfoliated graphite as an effective electrode material for supercapacitors. Surf. Eng. Appl. Electrochem. 2015, 51, 501–508. [Google Scholar] [CrossRef]
  11. Bannov, A.G.; Yusin, S.I.; Timofeeva, A.A.; Dyukova, K.D.; Ukhina, A.V.; Maksimovskii, E.A.; Popov, M.V. Synthesis of exfoliated graphite and its use as an electrode in supercapacitors. Prot. Met. Phys. Chem. Surf. 2016, 52, 645–652. [Google Scholar] [CrossRef]
  12. Dong, X.; Chen, H.; Lai, H.; Wang, L.; Wang, J.; Fang, W.; Gao, C. A graphitized expanded graphite cathode for aluminum-ion battery with excellent rate capability. J. Energy Chem. 2022, 66, 38–44. [Google Scholar] [CrossRef]
  13. Hou, S.; Li, J.; Huang, X.; Wang, X.; Ma, L.; Shen, W.; Kang, F.; Huang, Z.-H. Silver nanoparticles-loaded exfoliated graphite and its anti-bacterial performance. Appl. Sci. 2017, 7, 852. [Google Scholar] [CrossRef] [Green Version]
  14. Mangu, R.; Rajaputra, S.; Singh, V.P. MWCNT-polymer composites as highly sensitive and selective room temperature gas sensors. Nanotechnology 2011, 22, 215502. [Google Scholar] [CrossRef]
  15. Wu, K.-H.; Huang, W.-C.; Hung, W.-C.; Tsai, C.-W. Modified expanded graphite/Fe3O4 composite as an adsorbent of methylene blue: Adsorption kinetics and isotherms. Mater. Sci. Eng. B 2021, 266, 115068. [Google Scholar] [CrossRef]
  16. Chen, X.; Li, H.; Zeng, T.; Zhang, Y.; Wan, Q.; Li, Y.; Yang, N. Three-dimensional catalyst systems from expanded graphite and metal nanoparticles for electrocatalytic oxidation of liquid fuels. Nanoscale 2019, 11, 7952–7958. [Google Scholar] [CrossRef]
  17. Saidaminov, M.I.; Maksimova, N.V.; Zatonskih, P.V.; Komarov, A.D.; Lutfullin, M.A.; Sorokina, N.E.; Avdeev, V.V. Thermal decomposition of graphite nitrate. Carbon 2013, 59, 337–343. [Google Scholar] [CrossRef]
  18. Yin, G.; Sun, Z.; Gao, Y.; Xu, S. Preparation of expanded graphite for malachite green dye removal from aqueous solution. Microchem. J. 2021, 166, 106190. [Google Scholar] [CrossRef]
  19. Jin, S.; Xie, L.S.; Ma, Y.L.; Han, J.J.; Xia, Z.; Zhang, G.X.; Dong, S.M.; Wang, Y.Y. Low-temperature expanded graphite for preparation of graphene sheets by liquid-phase method. J. Phys. Conf. Ser. 2009, 188, 012040. [Google Scholar] [CrossRef]
  20. Yakovlev, A.V.; Finaenov, A.I.; Zabud’kov, S.L.; Yakovleva, E.V. Thermally expanded graphite: Synthesis, properties, and prospects for use. Russ. J. Appl. Chem. 2006, 79, 1741–1751. [Google Scholar] [CrossRef]
  21. Hong, X.; Chung, D.D.L. Exfoliated graphite with relative dielectric constant reaching 360, obtained by exfoliation of acid- intercalated graphite flakes without subsequent removal of the residual acidity. Carbon 2015, 91, 1–10. [Google Scholar] [CrossRef]
  22. Afanasov, I.M.; Shornikova, O.N.; Vlasov, I.I.; Kogan, E.V.; Seleznev, A.N.; Avdeev, V.V. Porous carbon materials based on exfoliated graphite. Inorg. Mater. 2009, 45, 135–139. [Google Scholar] [CrossRef]
  23. Li, M.; Liu, J.; Pan, S.; Zhang, J.; Liu, Y.; Liu, J.; Lu, H. Highly Oriented Graphite Aerogel Fabricated by Confined Liquid-Phase Expansion for Anisotropically Thermally Conductive Epoxy Composites. ACS Appl. Mater. Interfaces 2020, 12, 27476–27484. [Google Scholar] [CrossRef] [PubMed]
  24. Van Heerden, X.; Badenhorst, H. The influence of three different intercalation techniques on the microstructure of exfoliated graphite. Carbon 2015, 88, 173–184. [Google Scholar] [CrossRef] [Green Version]
  25. Yu, Z.; Nan, F.; Su, L.; Zhang, S.; He, Y. Effect of Ammonium Bicarbonate on Intercalation and Exfoliation of Graphite Materials. J. Nanomater. 2019, 2019, 1–8. [Google Scholar] [CrossRef]
  26. Chung, D.D.L. A review of exfoliated graphite. J. Mater. Sci. 2015, 51, 554–568. [Google Scholar] [CrossRef]
  27. Wang, A.; Chung, D.D.L. Dielectric and electrical conduction behavior of carbon paste electrochemical electrodes, with decoupling of carbon, electrolyte and interface contributions. Carbon 2014, 72, 135–151. [Google Scholar] [CrossRef]
  28. Goudarzi, R.; Motlagh, G.H. The effect of graphite intercalated compound particle size and exfoliation temperature on porosity and macromolecular diffusion in expanded graphite. Heliyon 2019, 5, e02595. [Google Scholar] [CrossRef] [Green Version]
  29. Asghar, H.M.A.; Hussain, S.N.; Sattar, H.; Brown, N.W.; Roberts, E.P.L. Environmentally friendly preparation of exfoliated graphite. J. Ind. Eng. Chem. 2014, 20, 1936–1941. [Google Scholar] [CrossRef]
  30. Asghar, H.M.A.; Hussain, S.N.; Sattar, H.; Brown, N.W.; Roberts, E.P.L. Potential Graphite Materials for the Synthesis of GICs. Chem. Eng. Commun. 2015, 202, 508–512. [Google Scholar] [CrossRef]
  31. Seredych, M.; Tamashausky, A.V.; Bandosz, T.J. Surface features of exfoliated graphite/bentonite composites and their importance for ammonia adsorption. Carbon 2008, 46, 1241–1252. [Google Scholar] [CrossRef]
  32. Manoj, B.; Kunjomana, A.G. Study of stacking structure of amorphous carbon by X-ray diffraction technique. Int. J. Electrochem. Sci. 2012, 7, 3127–3134. [Google Scholar]
  33. Bannov, A.G.; Uvarov, N.F.; Ukhina, A.V.; Chukanov, I.S.; Dyukova, K.D.; Kuvshinov, G.G. Structural changes in carbon nanofibers induced by ball milling. Carbon 2012, 50, 1090–1098. [Google Scholar] [CrossRef]
  34. Shiraishi, M.; Inagaki, M. Chapter 10—X-ray Diffraction Methods to Study Crystallite Size and Lattice Constants of Carbon Materials. In Carbon Alloys; Yasuda, E., Inagaki, M., Kaneko, K., Endo, M., Oya, A., Tanabe, Y.B.T.-C.A., Eds.; Elsevier Science: Oxford, UK, 2003; pp. 161–173. ISBN 978-0-08-044163-4. [Google Scholar]
  35. Maire, J.; Mering, J. Chemistry and Physics of Carbon; CRC Press: Boca Raton, FL, USA, 1975. [Google Scholar]
  36. Bannov, A.G.; Popov, M.V.; Kurmashov, P.B. Thermal analysis of carbon nanomaterials: Advantages and problems of interpretation. J. Therm. Anal. Calorim. 2020, 142, 349–370. [Google Scholar] [CrossRef]
  37. Leshin, V.S.; Sorokina, N.E.; Avdeev, V.V. Electrochemical synthesis and thermal properties of graphite bisulfate. Inorg. Mater. 2004, 40, 649–655. [Google Scholar] [CrossRef]
  38. Huang, J.; Tang, Q.; Liao, W.; Wang, G.; Wei, W.; Li, C. Green Preparation of Expandable Graphite and Its Application in Flame-Resistance Polymer Elastomer. Ind. Eng. Chem. Res. 2017, 56, 5253–5261. [Google Scholar] [CrossRef]
  39. Hong, Y.; Wang, Z.; Jin, X. Sulfuric acid intercalated graphite oxide for graphene preparation. Sci. Rep. 2013, 3, 1–6. [Google Scholar] [CrossRef]
  40. Xu, C.; Jiao, C.; Yao, R.; Lin, A.; Jiao, W. Adsorption and regeneration of expanded graphite modified by CTAB-KBr/H3PO4 for marine oil pollution. Environ. Pollut. 2018, 233, 194–200. [Google Scholar] [CrossRef]
  41. Li, W.; Han, C.; Liu, W.; Zhang, M.; Tao, K. Expanded graphite applied in the catalytic process as a catalyst support. Catal. Today 2007, 125, 278–281. [Google Scholar] [CrossRef]
  42. Lan, R.; Su, W.; Li, J. Preparation and catalytic performance of expanded graphite for oxidation of organic pollutant. Catalysts 2019, 9, 280. [Google Scholar] [CrossRef] [Green Version]
  43. Salvatore, M.; Carotenuto, G.; De Nicola, S.; Camerlingo, C.; Ambrogi, V.; Carfagna, C. Synthesis and Characterization of Highly Intercalated Graphite Bisulfate. Nanoscale Res. Lett. 2017, 12, 1–8. [Google Scholar] [CrossRef] [Green Version]
  44. Sing, K.S.W.; Williams, R.T. Physisorption hysteresis loops and the characterization of nanoporous materials. Adsorpt. Sci. Technol. 2004, 22, 773–782. [Google Scholar] [CrossRef]
  45. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  46. Zaaba, N.I.; Foo, K.L.; Hashim, U.; Tan, S.J.; Liu, W.W.; Voon, C.H. Synthesis of Graphene Oxide using Modified Hummers Method: Solvent Influence. Procedia Eng. 2017, 184, 469–477. [Google Scholar] [CrossRef]
  47. Greenwood, N.N. Spectroscopic Properties of Inorganic and Organometallic Compounds; Royal Society of Chemistry: London, UK, 2007; Volume 5, p. 630. [Google Scholar] [CrossRef]
  48. Hao, R.; Qian, W.; Zhang, L.; Hou, Y. Aqueous dispersions of TCNQ-anion-stabilized graphene sheets. Chem. Commun. 2008, 48, 6576–6578. [Google Scholar] [CrossRef] [PubMed]
  49. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved synthesis of graphene oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  50. Seredych, M.; Tamashausky, A.V.; Bandosz, T.J. Graphite oxides obtained from porous graphite: The role of surface chemistry and texture in ammonia retention at ambient conditions. Adv. Funct. Mater. 2010, 20, 1670–1679. [Google Scholar] [CrossRef]
  51. Justh, N.; Berke, B.; László, K.; Szilágyi, I.M. Thermal analysis of the improved Hummers’ synthesis of graphene oxide. J. Therm. Anal. Calorim. 2018, 131, 2267–2272. [Google Scholar] [CrossRef] [Green Version]
  52. Liao, N.; Li, Y.; Shan, J.; Zhu, T.; Sang, S.; Jia, D. Improved oxidation resistance of expanded graphite through nano SiC coating. Ceram. Int. 2018, 44, 3319–3325. [Google Scholar] [CrossRef]
  53. Saidaminov, M.I.; Maksimova, N.V.; Avdeev, V.V. Expandable graphite modification by boric acid. J. Mater. Res. 2012, 27, 1054–1059. [Google Scholar] [CrossRef]
  54. Savchenko, D.V.; Serdan, A.A.; Morozov, V.A.; van Tendeloo, G.; Ionov, S.G. Improvement of the oxidation stability and the mechanical properties of flexible graphite foil by boron oxide impregnation. New Carbon Mater. 2012, 27, 12–18. [Google Scholar] [CrossRef]
  55. Qiu, Y.; Moore, S.; Hurt, R.; Külaots, I. Influence of external heating rate on the structure and porosity of thermally exfoliated graphite oxide. Carbon 2017, 111, 651–657. [Google Scholar] [CrossRef] [Green Version]
  56. Stankovich, S.; Dikin, D.; Piner, R.D.; Kohlhaas, K.; Kleinhammes, A.; Jia, Y.; Wu, Y.; Nguyen, S.T.; Ruoff, R.S. Synthesis of graphene-based nanosheets via chemical reduction of exfoliated graphite oxide. Carbon 2007, 45, 1558–1565. [Google Scholar] [CrossRef]
  57. Kang, F.; Zheng, Y.-P.; Wang, H.-N.; Nishi, Y.; Inagaki, M. Effect of preparation conditions on the characteristics of exfoliated graphite. Carbon 2002, 40, 1575–1581. [Google Scholar] [CrossRef]
  58. Tichapondwa, S.M.; Tshemese, S.; Mhike, W. Adsorption of phenol and chromium (VI) pollutants in wastewater using exfoliated graphite. Chem. Eng. Trans. 2018, 70, 847–852. [Google Scholar] [CrossRef]
  59. Krawczyk, P.; Gurzęda, B. Electrochemical properties of exfoliated graphite affected by its two-step modification. J. Solid State Electrochem. 2016, 20, 361–369. [Google Scholar] [CrossRef]
Figure 1. Appearance of the TEG1-400 (a) and TEG2-400 (b) samples.
Figure 1. Appearance of the TEG1-400 (a) and TEG2-400 (b) samples.
Materials 14 07687 g001
Figure 2. TG/DSC curves of intercalated graphite (Ar, 2.5 °C/min).
Figure 2. TG/DSC curves of intercalated graphite (Ar, 2.5 °C/min).
Materials 14 07687 g002
Figure 3. SEM images of the samples: (a,b)–intercalated graphite; (c,d)–TEG1-400 sample, obtained by programmable heating; (e,f)–TEG2-400 sample, obtained by thermal shock.
Figure 3. SEM images of the samples: (a,b)–intercalated graphite; (c,d)–TEG1-400 sample, obtained by programmable heating; (e,f)–TEG2-400 sample, obtained by thermal shock.
Materials 14 07687 g003
Figure 4. Dependence of bulk density of EG, synthesized by programmable heating, on the synthesis temperature.
Figure 4. Dependence of bulk density of EG, synthesized by programmable heating, on the synthesis temperature.
Materials 14 07687 g004
Figure 5. SEM images of the EG samples: TEG1-500 (a), TEG1-600 (b), TEG1-700 (c), TEG2-1000 (d).
Figure 5. SEM images of the EG samples: TEG1-500 (a), TEG1-600 (b), TEG1-700 (c), TEG2-1000 (d).
Materials 14 07687 g005
Figure 6. XRD patterns of the intercalated graphite and TEG1-400 (a) and 002 reflection of these samples (b).
Figure 6. XRD patterns of the intercalated graphite and TEG1-400 (a) and 002 reflection of these samples (b).
Materials 14 07687 g006
Figure 7. Link between specific surface area, number of carbon layers in crystallite and crystallite width Lc. (a), Pore size distribution curves of EG: TEG1-400 (b), TEG1-500 (c), TEG1-600 (d), TEG1-700 (e), TEG2-400 (f), TEG2-1000 (g).
Figure 7. Link between specific surface area, number of carbon layers in crystallite and crystallite width Lc. (a), Pore size distribution curves of EG: TEG1-400 (b), TEG1-500 (c), TEG1-600 (d), TEG1-700 (e), TEG2-400 (f), TEG2-1000 (g).
Materials 14 07687 g007
Figure 8. FTIR spectra of IG and TEG1-400.
Figure 8. FTIR spectra of IG and TEG1-400.
Materials 14 07687 g008
Figure 9. Raman spectra of samples synthesized by programmable heating (a) and thermal shock (b).
Figure 9. Raman spectra of samples synthesized by programmable heating (a) and thermal shock (b).
Materials 14 07687 g009
Figure 10. C1s XPS spectra of intercalated graphite (a) and TEG1-400 sample (b); O1s spectra of intercalated graphite and TEG1-400 sample (c).
Figure 10. C1s XPS spectra of intercalated graphite (a) and TEG1-400 sample (b); O1s spectra of intercalated graphite and TEG1-400 sample (c).
Materials 14 07687 g010
Figure 11. DSC curves of EG samples (10 K/min) (exo down).
Figure 11. DSC curves of EG samples (10 K/min) (exo down).
Materials 14 07687 g011
Table 1. Samples synthesized under different conditions.
Table 1. Samples synthesized under different conditions.
SampleTemperature, °CSynthesis Technique
TEG1-400400Programmable heating
TEG1-500500Programmable heating
TEG1-600600Programmable heating
TEG1-700700Programmable heating
TEG2-400400Thermal shock
TEG2-10001000Thermal shock
Table 2. Concentration of the main elements in the EG samples (EDX).
Table 2. Concentration of the main elements in the EG samples (EDX).
ElementConcentration, at.%
Intercalated GraphiteTEG1-400TEG1-500TEG1-600TEG1-700TEG2-400TEG2-1000
C87.3195.1396.4795.6997.3793.2697.46
O11.064.323.053.752.176.212.14
S1.140.290.280.230.140.370.05
Table 3. Yield, bulk density, and textural characteristics of samples synthesized by programmable heating and thermal shock.
Table 3. Yield, bulk density, and textural characteristics of samples synthesized by programmable heating and thermal shock.
SampleTotal Surface Area, m2/gSurface Area of Micropores, m2/gSurface Area of Mesopores, m2/gTotal Pore Volume, cm3/gAverage Pore Diameter, nmYield, %Bulk Density, g/cm3
TEG1-4006992614380.9455.489.20.058
TEG1-5006261634630.7364.779.00.018
TEG1-600257851720.3255.078.70.018
TEG1-700229951340.2314.077.80.017
TEG2-40018490940.1984.377.40.016
TEG2-10006392044350.7614.876.90.019
Table 4. Raman spectroscopy data of EG samples.
Table 4. Raman spectroscopy data of EG samples.
SampleD Peak Position, cm−1DFWHM, cm−1G Peak Position, cm−1GFWHM,
cm−1
I (D)/I (G)
TEG1-4001357531584200.1
TEG1-5001359191584170.0104
TEG1-6001355361584160.0125
TEG1-7001357331584180.038
TEG2-4001357221584170.014
TEG2-10001357221584150.007
Table 5. Analysis of DSC peak of oxidation of EG samples.
Table 5. Analysis of DSC peak of oxidation of EG samples.
SamplePeak Onset, Tonset (°C)Peak Maximum, Tmax (°C)Peak End, Tend (°C)Heat Release, J/g
TEG1-4006527408009582
TEG1-50067576481212,207
TEG1-60063974882912,975
TEG1-70063775382812,789
TEG2-40064277384111,967
TEG2-100064173380313,773
Table 6. Surface area values of expanded graphite and related materials reported in literature depending on precursor type, synthesis technique, and synthesis temperature.
Table 6. Surface area values of expanded graphite and related materials reported in literature depending on precursor type, synthesis technique, and synthesis temperature.
MaterialPrecursorSpecific Surface Area (BET), m2/gSynthesis TechniqueSynthesis Temperature, °CReference
Expanded graphiteIntercalated graphite678Programmable heating (20 °C/min)400This work
Expanded graphiteGraphite oxide (modified Hummers method)400Thermal shock (1570 °C/min)500–900[55]
Expanded graphiteGraphite bisulfate77Thermal shock1300[1]
Expanded graphiteGraphene oxide(Hummers method)466Exfoliation in aqueous median/a[56]
Expanded graphiteExpandable natural graphite flake (interca- lated by using sulfuric acid) from45Thermal shock900[21]
Expanded graphiteExpandable graphite (H2SO4–GIC)10–40Thermal shock400–1000[57]
Expanded graphiteExpandable graphite22.4Thermal shock600[58]
Sulfur-free Expanded graphiteGIC245Thermal shock950[9]
Re-expanded EGExpanded EG33.5Thermal shock800[59]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bannov, A.G.; Ukhina, A.V.; Maksimovskii, E.A.; Prosanov, I.Y.; Shestakov, A.A.; Lapekin, N.I.; Lazarenko, N.S.; Kurmashov, P.B.; Popov, M.V. Highly Porous Expanded Graphite: Thermal Shock vs. Programmable Heating. Materials 2021, 14, 7687. https://doi.org/10.3390/ma14247687

AMA Style

Bannov AG, Ukhina AV, Maksimovskii EA, Prosanov IY, Shestakov AA, Lapekin NI, Lazarenko NS, Kurmashov PB, Popov MV. Highly Porous Expanded Graphite: Thermal Shock vs. Programmable Heating. Materials. 2021; 14(24):7687. https://doi.org/10.3390/ma14247687

Chicago/Turabian Style

Bannov, Alexander G., Arina V. Ukhina, Evgenii A. Maksimovskii, Igor Yu. Prosanov, Artyom A. Shestakov, Nikita I. Lapekin, Nikita S. Lazarenko, Pavel B. Kurmashov, and Maksim V. Popov. 2021. "Highly Porous Expanded Graphite: Thermal Shock vs. Programmable Heating" Materials 14, no. 24: 7687. https://doi.org/10.3390/ma14247687

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