Next Article in Journal
Effects of Short Fibers on the Long-Term Behavior of RC/FRC Beams Aged under Service Loading
Next Article in Special Issue
High-Frequency Limits of Graphene Field-Effect Transistors with Velocity Saturation
Previous Article in Journal
Dynamic Modeling of McKibben Muscle Using Empirical Model and Particle Swarm Optimization Method
Previous Article in Special Issue
Nonlinear Refractive Index in Rectangular Graphene Quantum Dots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Preparation of Few-Layer Graphene Dispersions from Hydrothermally Expanded Graphite

1
Physics Research Group, Faculty of Science, Escuela Superior Politécnica de Chimborazo, Riobamba EC060101, Ecuador
2
GraphenTech NL, Olympiaweg 28A, 3077AL Rotterdam, The Netherlands
3
Faculty of Mechanical Engineering, Escuela Superior Politécnica de Chimborazo, Riobamba EC060101, Ecuador
4
Research Institute on Membrane Technology (ITM - CNR), Cubo 17C, 87036 University of Calabria, 87036 Cosenza, Italy
5
Grupo de Investigación en Ambiente y Desarrollo, Escuela Superior Politécnica de Chimborazo, Riobamba EC060101, Ecuador
6
INFN-Laboratori Nazionali di Frascati, Via E. Fermi 40, I-00044 Frascati, Italy
*
Author to whom correspondence should be addressed.
Appl. Sci. 2019, 9(12), 2539; https://doi.org/10.3390/app9122539
Submission received: 28 May 2019 / Revised: 14 June 2019 / Accepted: 18 June 2019 / Published: 21 June 2019
(This article belongs to the Special Issue Graphene and Graphene-Based Composites for Electronics)

Abstract

:
In this study, we propose a novel approach to prepare few-layer graphene (FLG) dispersions, which is realized by exfoliating natural graphite flakes in a surfactant aqueous solution under hydrothermal treatment and liquid-phase exfoliation. In order to obtain stable and well-dispersed FLG dispersions, pristine graphite is hydrothermally expanded in a hexadecyltrimethylammonium bromide (CTAB) aqueous solution at 180 °C for 15 h, followed by sonication up to 3 h. In comparison to long-time sonication methods, the present method is significantly efficient, and most importantly, does not involve the use of an oxidizing agent and hazardous media, which will make it more competent in the scalable production of graphene.

Graphical Abstract

1. Introduction

Owing to its outstanding physical and chemical properties [1,2], graphene has become the most studied nanomaterial of the past decade. Until now, many top-down methods have developed for preparing graphene, such as micromechanical cleavage of graphite [3], oxidation-reduction of graphite [4,5], electrochemical exfoliation of graphite [6], and liquid-phase exfoliation of graphite [7]. Among them, the liquid-phase exfoliation is the most commonly used method to prepare graphene by sonication at a low cost and with minimal environmental impact. However, this method mostly involves the use of high-boiling point solvents, e.g., N-methyl-pyrrolidone (NMP) [8] and a long sonication time to reach high concentrations of dispersed graphene [9].
A viable liquid-phase exfoliation route is carried out in an aqueous solution through suitable surfactants [10] (and biosurfactants [11]) that reduce and optimize the high surface energy of water to match with highly hydrophobic graphitic surfaces [12]. Surfactants offer the steric and electrostatic repulsion that is necessary to avoid the graphene sheets from aggregation due to the sonication which breaks up graphite into few-layer graphene (FLG) flakes that are coated by the surfactant molecules [12,13]. Surfactants, therefore, are excellent candidates in the direct exfoliation of graphite because they are used as a dispersant and a stabilizer.
In an effort to circumvent long sonication time limitations (e.g., the induced defects in graphene structure [14]), Ou et al. [15] successfully demonstrated the preparation of FLG dispersions from graphite powder, in a relatively short sonication time by combining heat-treatment in NMP and liquid-phase exfoliation in different organic solvents. The heat-treatment increases the interlayer distance of graphite, which is stabilized by adding polyvinyl pyrrolidone (PVP), followed by sonication for up to 20 h to prepare well-dispersed FLG dispersions. To our knowledge, an alternative method that combines heat-treatment of natural graphite flakes and direct exfoliation in stabilizing surfactants is unexplored, particularly, at a short sonication time.
In this study, we propose such a method to prepare FLG dispersions in surfactant water solution using hydrothermal treatment and liquid-phase exfoliation. First, we subject pristine natural graphite to temperature to produce pretreated graphite flakes. Subsequently, pretreated graphite is exfoliated into FLG by sonication and centrifugation. The cationic surfactant, hexadecyltrimethylammonium bromide (CTAB), is selected as a dispersant and a stabilizer [16,17]. CTAB is commonly considered as the dispersing agent to stabilize aqueous solutions of graphene [18] and it has also been used as the precursor for preparing graphene nanocomposites [19] with promising applications in electrochemical sensing and biosensing [20].

2. Materials and Methods

Natural graphite flakes were purchased from Pingdu Huadong Graphite Co. Ltd., and the CTAB (H5882-100G, 98%, M.W. 364.45) was purchased from Sigma Aldrich. The graphite and the CTAB were used as received, without further purification. In a typical experiment, 100 mg of natural graphite flakes were added into 50 mL of CTAB aqueous solution at different surfactant concentrations (i.e., 0.1, 0.5, and 1.0 mg mL−1). The resulting mixture was transferred into a sealed Teflon vessel (100 mL) and reacted at 180   ° C for 15 h (optimal parameters reported in [15,21]) to obtain pretreated graphite. After the hydrothermal treatment (HT), pretreated graphite was immediately dispersed in the very same CTAB aqueous solution by sonication, employing a tip sonicator (UP100H-Hielscher, 50 watts, 30 kHz) in continuous operation. To reduce the damage to the graphene sheets by ultrasonic processing, the maximum sonication time was 3 h. The sonication produced black suspensions that were left to stand overnight and then centrifugated for 90 min at 2000 rpm (optimal parameters reported in [22]) to remove nonexfoliated graphite flakes. After centrifugation, the pipette-extracted 40 mL supernatant was semitransparent without the presence of agglomerates (Supplementary Figure S1). The concentration of dispersed graphene was determined by UV-visible absorption spectroscopy and the absorbance ( A ) was adapted to calculate the concentration using the Beer–Lambert law (a widely used approach [10,22,23]):
C G = A l   ·   α 660
where, l = 0.01 m, is the path length and α = 1390 mL mg−1 m−1, is the absorption coefficient of the dispersed graphene in the surfactant water solutions [10]. The characteristic absorption spectra of the dispersed graphene were recorded using UV-vis spectroscopy (UV-160A, Shimadzu Corporation, Japan). The morphologies of the samples were observed using a scanning electron microscopy (SEM, Quanta Feg 400 F7) with accelerating voltage ranging from 5 to 35 kV. Raman spectra were obtained using a Jobin Yvon LABRAM spectrometer, with a 514 nm laser wavelength.

3. Results and Discussions

To begin, we briefly analyze the hydrothermal conditions, the sonication process, the surfactant effect, and the yield. Supplementary Figure S2 shows the schematic route of the preparation process for graphene dispersions. Under hydrothermal conditions, various factors affect the graphite crystal structure which include temperature and pressure. Temperature generates water vapor, and pressure origins changes in the kinetics of reactions affecting the graphite structure [24]. Therefore, temperature and pressure create conditions for the intercalation of the CTAB (and water) molecules between graphene layers [25]. The intercalation increases the interlayer distance of graphite which is crucial to the improvement of the subsequent ultrasonication process [15].
Ultrasonication is described by the effect of acoustic cavitation of high frequency ultrasound in the formation, growth, and collapse of microbubbles in solution, which induces shock waves on the graphite surface, producing exfoliation of graphite into FLG [26]. Notwithstanding, the direct exfoliation of graphite in water is possible and the resulting exfoliated graphene flakes are not stable. The CTAB molecules have a cationic ammonium head group and an extended alkyl chain as a tail [26]. It is expected that the CTAB molecules are adsorbed onto the graphene surface through electrostatic interactions [25]. The hydrophobic interaction of the alkyl chains of the CTAB molecules on the graphene surface prevents the re-staking and agglomeration of exfoliated flakes.
The yield of this approach, based on the absorption coefficient of [10], was 5 % and needs to be optimized. At this point, the apparent low concentration of dispersed graphene in comparison with the relative amount of CTAB, is the result of using the absorbance to calculate the concentration by the Beer–Lambert law. Since this approach is usually used to obtain the concentration of dispersed graphene [10,22,23], it is employed to emphasize the role of hydrothermal treatment instead of its final yield. Attempts to improve the yield are currently in progress using the gravimetric method and extending the study for cationic/anionic surfactants.
Pristine (gray) graphite and pretreated (light-yellow) graphite are shown in Figure 1a,b, respectively. The optical change is due to the interaction of graphite with the surfactant, and therefore the obtained light-yellow graphite is due to the presence of molecular bromine because the resulting pretreated graphite was not washed after HT, which is verified by the surfactant bubbles in Figure 1b. Since pretreated graphite is dispersed in the very same CTAB aqueous solution, sonication is immediately used to simplify the method.
In order to understand the role of HT, Figure 1a,b present the concentration of dispersed graphene considering different amounts of CTAB and different sonication times. Again, we stress that this study is limited to 3 h of sonication, however it is well-known that the exfoliation yield is increased by increasing the sonication time, and a suitable sonication time is 10 h [15]. We show, however, that the concentration of dispersed graphene from pretreated graphite is significantly enhanced, suggesting that HT plays a crucial role. Therefore, depending on the amount of CTAB, sonication time, and HT, the concentration of dispersed graphene ranges from 40 μg mL−1 to 60 μg mL−1. These concentrations are comparable to those obtained in early studies by exfoliating graphite with a long sonication time [22,27].
CTAB seems to be an efficient dispersant at high concentrations (0.5, 1.0 mg mL−1). However, high quantities of surfactant may have a significant impact on the electrical properties of graphene because the residual surfactant is difficult to remove [28]. To circumvent this, we focus on graphene dispersions obtained at the lowest concentration of CTAB (0.1 mg mL−1) and after 3 h of sonication. Figure 1c shows the stability of graphene dispersions, at the lowest concentration of CTAB, obtained from pretreated and pristine graphite. In both cases, graphene flakes are easy to precipitate, as the storage time increases, however, it is noteworthy that the stability of graphene dispersion obtained from pretreated graphite is higher than from pristine graphite within 12 weeks. These results corroborate the importance of HT. The UV-visible absorption spectra of graphene dispersions, after 3 h of sonication and different CTAB concentrations, are shown in Figure 1d. The spectra are featureless in the visible region as expected for graphene. A prominent peak is found at 266 nm, corresponding to the π to π * transitions of graphene and graphite [29].
Previously, the transformation of the graphite (powder) structure, after heat-treatment in NMP, was effectively demonstrated by SEM, Raman, and X-ray diffraction (XRD) measurements, i.e., the increased interlayer spacing of graphite as a direct consequence of increasing the heat [15]. In Figure 2, we confirm this fact by scrutinizing the morphology of pretreated (natural) graphite flakes considering the different amounts of CTAB.
Independently of the CTAB concentration (0.1, 0.5 and 1.0 mg mL−1), as shown in Figure 2a–c, HT also increases the interlayer spacing of graphite which is stabilized by surfactant. Therefore, the resulting pretreated graphite is called hydrothermally expanded graphite (HEG). At the highest concentration of CTAB, in Figure 2a, we show the surfactant intercalation (green arrows) and the HEG surface coated by the surfactant molecules (green dashed circle). After 3 h of sonication in the three different CTAB concentrations, Figure 2d–f demonstrate that graphene dispersions are composed of smaller and thinner layers.
These layers seem to be semitransparent and surfactant-free, which is interpreted as an important result for future applications in electronic devices. However, by scrutinizing the morphology of layers at the highest concentration of CTAB and by changing the accelerating voltage from 5 to 30 kV, in Supplementary Figure S3, we observe the presence of surfactant which corroborates the statement about the difficulty of the residual surfactant removal and the need to use the proposed method at the lowest concentration of CTAB in order to prevent the intrinsic properties of the obtained layers.
The Raman spectrum of graphite and dispersed graphene (at the lowest concentration of CTAB, after HT and 3 h of sonication) are reported in Figure 3. We detect the three major peaks, i.e., the D peak at 1340   cm−1, the G peak at 1580 cm−1 ( 1572 cm−1), and the 2D peak at 2719 cm−1 ( 2711 cm−1). The D peak and G peak are attributed to the edge/basal defects and the ordered sp2-hybridized carbon bonds in the graphene/graphite lattice, respectively. In dispersed graphene, the peak ratio between the intensity of D and G peaks (ID/IG) is 0.28, indicating that some edge/basal defects are induced during the process of ultrasonication [30]. Probably, the small size and folded edge of graphene layers (Figure 2d–f), contributes significantly to the observed D peak, although we cannot completely rule out the presence of in-plane defects.
The structure of G peak observed in the dispersed graphene at 1572   cm−1 appears to be slightly affected by the presence of D’ peak, however, the D’ peak usually is present when there are surface defects, such as charging or other impurities adsorbed onto the surface [31]. The latter here is attributed to the remaining surfactant as observed in Supplementary Figure S3 at the highest concentration of CTAB, which strongly confirms the need to use our method at low concentrations of surfactant. Interestingly enough, the D and G (G and D’) peaks of our graphene sample are well separated from each other and are distinctly different from those of graphene oxide, in which case the D and G peaks are broad and overlap. This fact suggests that our dispersed graphene did not undergo severe in-plane disruption as in the case of graphene oxide [30].
The peak is slightly broader between 2625   and   2700 cm−1 and it is commonly used to estimate the number of layers in obtained graphene [32]. However, the intensity of the 2D peak depends on the excitation laser frequency, and therefore cannot be solely relied upon [31]. In this context, we use the full width at half maximum (FWHM) to evaluate the 2D peak, which is found to be a qualitative guide to distinguish the number of layers from single- to five-layer graphene [33]. By fitting the 2D peak with two Lorentzian functions, the intensity of the 2D1A peak (detected at 2683 cm−1) increases in dispersed graphene, with respect to the intensity of the 2D1A peak (at 2676 cm−1) in pristine graphite. Most importantly, in Supplementary Figures S4 and S5, we report the FWHM of 2D1A =   68.3   ±   1.6 cm−1, 2D2A =   36.9   ±   1.0 cm−1, and 2D =   68.7   ±   1.1 cm−1, respectively. These values are in good agreement with previous reports for tri-, four-, and five-layer graphene [33]. Furthermore, the 2D peak became more symmetrical in dispersed graphene, after HT and sonication, indicating the presence of FLG in obtained graphene dispersions.

4. Conclusions

In this study, a simple method to prepare graphene dispersions has been demonstrated. Natural graphite flakes were effectively exfoliated into FLG under HT and liquid-phase exfoliation in a surfactant aqueous solution. The cationic surfactant, CTAB, was used as a dispersant and a stabilizer. First, pristine graphite was hydrothermally expanded in the CTAB aqueous solution reacted at 180   ° C for 15 h. Subsequently, pretreated graphite (denominated as hydrothermally expanded graphite, HEG) was exfoliated into FLG by sonication for 3 h, producing stable and well-dispersed graphene dispersions. The expansion of interlayer spacing of graphite was demonstrated using SEM measurements. The efficiency of HT was estimated by measuring the concentration of the dispersed graphene which was shown to be comparable to graphene dispersions reported with a long sonication time [22,27]. Depending on the amount of the CTAB, sonication time, and HT, the concentration of dispersed graphene ranged from 40 μg mL−1 to 60 μg mL−1. The obtained graphene dispersions from pretreated graphite were shown to have good stability within 12 weeks. Using Raman analyses, the obtained FLG presented lower (edge/basal) defects, and most importantly, based on the FWHM analysis, we confirmed the presence of FLG, likely, no more than five layers. This novel approach (but non-optimized) may be explored in the future for cationic/anionic surfactants (or biosurfactants) in order to improve its efficiency.

Supplementary Materials

The following are available online at https://www.mdpi.com/2076-3417/9/12/2539/s1, Figure S1: Obtained graphene dispersions (12 weeks stored) after hydrothermal treatment (HT), 3 h of sonication and centrifugation, in the three different concentrations of CTAB (0.1, 0.5, and 1.0 mg mL−1) reported in the main text. After sonication, the obtained suspensions were left to stand overnight and then centrifugated for 90 min at 2000 rpm., Figure S2: Schematic route of the preparation process for FLG dispersions. Natural graphite flakes are added into CTAB aqueous solution (at different surfactant concentrations: 0.1, 0.5, and 1.0 mg mL−1). The resulting mixture is reacted at 180 °C for 15 h. The obtained pretreated graphite is immediately dispersed in the very same CTAB aqueous solution by sonication, Figure S3: Scanning electron microscopy (SEM) morphology of dispersed graphene layer in 1.0 mg mL−1 of CTAB aqueous solution, by changing the accelerating voltage from 5 to 30 kV. Semitransparent layer extracted from supernatant after HT, 3 h of sonication, centrifugation for 90 min at 2000 rpm, and dried at 235 °C, Figure S4: Raman spectrum of the 2D peak of dispersed graphene in 0.1 mg mL−1 of CTAB aqueous solution and fitting of the 2D peaks with two Lorentzian functions. The w values represent the respective full width at half maximum (FWHM) values, Figure S5: Raman spectrum of the 2D peak of dispersed graphene in 0.1 mg mL−1 of CTAB aqueous solution and fitting of the 2D peaks with one Lorentzian function. The w value represents the respective FWHM value.

Author Contributions

All authors contributed to designing and performing measurements, data analysis, scientific discussions, and writing the article.

Funding

This research received no external funding.

Acknowledgments

We acknowledge E. Cazzanelli for helping us with Raman measurements, as well as L. Caputi, Diana Coello-Fiallos, and Richard Pachacama for the stimulating discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Coello-Fiallos, D.; Tene, T.; Guayllas, J.L.; Haro, D.; Haro, A.; Gomez, C.V. DFT comparison of structural and electronic properties of graphene and germanene: Monolayer and bilayer systems. Mater. Today Proc. 2017, 4, 6835–6841. [Google Scholar] [CrossRef]
  2. Dreyer, D.R.; Park, S.; Bielawski, C.W.; Ruoff, R.S. The chemistry of graphene oxide. Chem. Soc. Rev. 2010, 39, 228–240. [Google Scholar] [CrossRef] [PubMed]
  3. Yi, M.; Shen, Z. A review on mechanical exfoliation for the scalable production of graphene. J. Mater. Chem. A 2015, 3, 11700–11715. [Google Scholar] [CrossRef]
  4. Chen, W.; Yan, L.; Bangal, P.R. Preparation of graphene by the rapid and mild thermal reduction of graphene oxide induced by microwaves. Carbon 2010, 48, 1146–1152. [Google Scholar] [CrossRef]
  5. Gomez, C.V.; Robalino, E.; Haro, D.; Tene, T.; Escudero, P.; Haro, A.; Orbe, J. Structural and electronic properties of graphene oxide for different degree of oxidation. Mater. Today Proc. 2016, 3, 796–802. [Google Scholar] [CrossRef]
  6. Parvez, K.; Li, R.; Puniredd, S.R.; Hernandez, Y.; Hinkel, F.; Wang, S.; Müllen, K. Electrochemically exfoliated graphene as solution-processable, highly conductive electrodes for organic electronics. ACS Nano 2013, 7, 3598–3606. [Google Scholar] [CrossRef] [PubMed]
  7. Hernandez, Y.; Nicolosi, V.; Lotya, M.; Blighe, F.M.; Sun, Z.; De, S.; Boland, J.J. High-yield production of graphene by liquid-phase exfoliation of graphite. Nat. Nanotechnol. 2008, 3, 563. [Google Scholar] [CrossRef]
  8. Usca, G.T.; Hernandez-Ambato, J.; Pace, C.; Caputi, L.S.; Tavolaro, A. Liquid-phase exfoliated graphene self-assembled films: Low-frequency noise and thermal-electric characterization. Appl. Surf. Sci. 2016, 380, 268–273. [Google Scholar] [CrossRef]
  9. Khan, U.; O’Neill, A.; Lotya, M.; De, S.; Coleman, J.N. High-concentration solvent exfoliation of graphene. Small 2010, 6, 864–871. [Google Scholar] [CrossRef]
  10. Lotya, M.; Hernandez, Y.; King, P.J.; Smith, R.J.; Nicolosi, V.; Karlsson, L.S.; Duesberg, G.S. Liquid phase production of graphene by exfoliation of graphite in surfactant/water solutions. J. Am. Chem. Soc. 2009, 131, 3611–3620. [Google Scholar] [CrossRef]
  11. Ba, H.; Truong-Phuoc, L.; Pham-Huu, C.; Luo, W.; Baaziz, W.; Romero, T.; Janowska, I. Colloid Approach to the Sustainable Top-Down Synthesis of Layered Materials. ACS Omega 2017, 2, 8610–8617. [Google Scholar] [CrossRef]
  12. Narayan, R.; Kim, S.O. Surfactant mediated liquid phase exfoliation of graphene. Nano Converg. 2015, 2, 20. [Google Scholar] [CrossRef] [PubMed]
  13. Smith, R.J.; Lotya, M.; Coleman, J.N. The importance of repulsive potential barriers for the dispersion of graphene using surfactants. New J. Phys. 2010, 12, 125008. [Google Scholar] [CrossRef]
  14. Ciesielski, A.; Samorì, P. Graphene via sonication assisted liquid-phase exfoliation. Chem. Soc. Rev. 2014, 43, 381–398. [Google Scholar] [CrossRef] [PubMed]
  15. Ou, E.; Xie, Y.; Peng, C.; Song, Y.; Peng, H.; Xiong, Y.; Xu, W. High concentration and stable few-layer graphene dispersions prepared by the exfoliation of graphite in different organic solvents. RSC Adv. 2013, 3, 9490–9499. [Google Scholar] [CrossRef]
  16. Wang, B.; Liang, X. Surface Decoration and Dispersibility of Graphene Nanoplatelets in Aqueous Surfactant Solution. J. Nanosci. Nanotechnol. 2019, 19, 2060–2069. [Google Scholar] [CrossRef]
  17. Vaghri, E.; Dorranian, D.; Ghoranneviss, M. Effects of CTAB concentration on the quality of graphene oxide nanosheets produced by green laser ablation. Mater. Chem. Phys. 2018, 203, 235–242. [Google Scholar] [CrossRef]
  18. Du, W.; Jiang, X.; Zhu, L. From graphite to graphene: Direct liquid-phase exfoliation of graphite to produce single-and few-layered pristine graphene. J. Mater. Chem. A 2013, 1, 10592–10606. [Google Scholar] [CrossRef]
  19. Wang, Z.; Chen, T.; Chen, W.; Chang, K.; Ma, L.; Huang, G.; Lee, J.Y. CTAB-assisted synthesis of single-layer MoS 2–graphene composites as anode materials of Li-ion batteries. J. Mater. Chem. A 2013, 1, 2202–2210. [Google Scholar] [CrossRef]
  20. Yang, Y.J.; Li, W. CTAB functionalized graphene oxide/multiwalled carbon nanotube composite modified electrode for the simultaneous determination of ascorbic acid, dopamine, uric acid and nitrite. Biosens. Bioelectron. 2014, 56, 300–306. [Google Scholar] [CrossRef]
  21. Zhong, C.; Wang, J.Z.; Wexler, D.; Liu, H.K. Microwave autoclave synthesized multi-layer graphene/single-walled carbon nanotube composites for free-standing lithium-ion battery anodes. Carbon 2014, 66, 637–645. [Google Scholar] [CrossRef] [Green Version]
  22. Lotya, M.; King, P.J.; Khan, U.; De, S.; Coleman, J.N. High-concentration, surfactant-stabilized graphene dispersions. ACS Nano 2010, 4, 3155–3162. [Google Scholar] [CrossRef] [PubMed]
  23. Ma, H.; Shen, Z.; Yi, M.; Ben, S.; Liang, S.; Liu, L.; Ma, S. Direct exfoliation of graphite in water with addition of ammonia solution. J. Colloid Interface Sci. 2017, 503, 68–75. [Google Scholar] [CrossRef] [PubMed]
  24. Ren, L.; Qi, X.; Liu, Y.; Hao, G.; Huang, Z.; Zou, X.; Zhong, J. Large-scale production of ultrathin topological insulator bismuth telluride nanosheets by a hydrothermal intercalation and exfoliation route. J. Mater. Chem. 2012, 22, 4921–4926. [Google Scholar] [CrossRef]
  25. Zhang, K.; Mao, L.; Zhang, L.L.; Chan, H.S.O.; Zhao, X.S.; Wu, J. Surfactant-intercalated, chemically reduced graphene oxide for high performance supercapacitor electrodes. J. Mater. Chem. 2011, 21, 7302–7307. [Google Scholar] [CrossRef]
  26. Vadukumpully, S.; Paul, J.; Valiyaveettil, S. Cationic surfactant mediated exfoliation of graphite into graphene flakes. Carbon 2009, 47, 3288–3294. [Google Scholar] [CrossRef]
  27. O’Neill, A.; Khan, U.; Nirmalraj, P.N.; Boland, J.; Coleman, J.N. Graphene dispersion and exfoliation in low boiling point solvents. J. Phys. Chem. C 2011, 115, 5422–5428. [Google Scholar] [CrossRef]
  28. Sukumaran, S.S.; Tripathi, S.; Resmi, A.N.; Gopchandran, K.G.; Jinesh, K.B. Influence of surfactants on the electronic properties of liquid-phase exfoliated graphene. Mater. Sci. Eng. B 2019, 240, 62–68. [Google Scholar] [CrossRef]
  29. Li, D.; Müller, M.B.; Gilje, S.; Kaner, R.B.; Wallace, G.G. Processable aqueous dispersions of graphene nanosheets. Nat. Nanotechnol. 2008, 3, 101. [Google Scholar] [CrossRef]
  30. Ricardo, K.B.; Sendecki, A.; Liu, H. Surfactant-free exfoliation of graphite in aqueous solutions. Chem. Commun. 2014, 50, 2751–2754. [Google Scholar] [CrossRef]
  31. Walch, N.J.; Nabok, A.; Davis, F.; Higson, S.P. Characterisation of thin films of graphene–surfactant composites produced through a novel semi-automated method. Beilstein J. Nanotechnol. 2016, 7, 209–219. [Google Scholar] [CrossRef] [PubMed]
  32. Ferrari, A.C.; Meyer, J.C.; Scardaci, V.; Casiraghi, C.; Lazzeri, M.; Mauri, F.; Geim, A.K. Raman spectrum of graphene and graphene layers. Phys. Rev. Lett. 2006, 97, 187401. [Google Scholar] [CrossRef] [PubMed]
  33. Hao, Y.; Wang, Y.; Wang, L.; Ni, Z.; Wang, Z.; Wang, R.; Thong, J.T. Probing layer number and stacking order of few-layer graphene by Raman spectroscopy. Small 2010, 6, 195–200. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Concentration of dispersed graphene as a function of three different hexadecyltrimethylammonium bromide (CTAB) concentrations (0.1, 0.5, 1.0 mg/mL−1) at three different sonication times, from (a) pristine (untreated) graphite and (b) treated graphite by the hydrothermal treatment (HT). (c) Stability of the graphene dispersion in 0.1 mg/mL−1 of the CTAB aqueous solution as a function of the storage time, considering pristine and pretreated graphite source. (d) UV-visible spectra of exfoliated graphene in the three different CTAB concentrations after HT and 3 h of sonication. Optical photos of pristine graphite (a, inset) and pretreated graphite (b, inset).
Figure 1. Concentration of dispersed graphene as a function of three different hexadecyltrimethylammonium bromide (CTAB) concentrations (0.1, 0.5, 1.0 mg/mL−1) at three different sonication times, from (a) pristine (untreated) graphite and (b) treated graphite by the hydrothermal treatment (HT). (c) Stability of the graphene dispersion in 0.1 mg/mL−1 of the CTAB aqueous solution as a function of the storage time, considering pristine and pretreated graphite source. (d) UV-visible spectra of exfoliated graphene in the three different CTAB concentrations after HT and 3 h of sonication. Optical photos of pristine graphite (a, inset) and pretreated graphite (b, inset).
Applsci 09 02539 g001
Figure 2. Scanning electron microscopy (SEM) morphology of pretreated natural graphite in (a) 1.0 mg mL−1, (b) 0.5 mg mL−1, and (c) 0.1 mg mL−1 of CTAB aqueous solution. (df) Semitransparent layers obtained from supernatant after 2000 rpm centrifugation for 90 min, subject to 3 h of sonication in the three different CTAB concentrations.
Figure 2. Scanning electron microscopy (SEM) morphology of pretreated natural graphite in (a) 1.0 mg mL−1, (b) 0.5 mg mL−1, and (c) 0.1 mg mL−1 of CTAB aqueous solution. (df) Semitransparent layers obtained from supernatant after 2000 rpm centrifugation for 90 min, subject to 3 h of sonication in the three different CTAB concentrations.
Applsci 09 02539 g002
Figure 3. Raman spectra of (a) dispersed graphene after hydrothermal treatment and 3 h of sonication in 0.1 mg mL−1 CTAB aqueous solution, fabricated by drop casting of dispersion, and (c) pristine graphite. (bd) Fitting of the 2D peaks with two Lorentzian functions. The intensity was normalized by the G peak.
Figure 3. Raman spectra of (a) dispersed graphene after hydrothermal treatment and 3 h of sonication in 0.1 mg mL−1 CTAB aqueous solution, fabricated by drop casting of dispersion, and (c) pristine graphite. (bd) Fitting of the 2D peaks with two Lorentzian functions. The intensity was normalized by the G peak.
Applsci 09 02539 g003

Share and Cite

MDPI and ACS Style

Vacacela Gomez, C.; Tene, T.; Guevara, M.; Tubon Usca, G.; Colcha, D.; Brito, H.; Molina, R.; Bellucci, S.; Tavolaro, A. Preparation of Few-Layer Graphene Dispersions from Hydrothermally Expanded Graphite. Appl. Sci. 2019, 9, 2539. https://doi.org/10.3390/app9122539

AMA Style

Vacacela Gomez C, Tene T, Guevara M, Tubon Usca G, Colcha D, Brito H, Molina R, Bellucci S, Tavolaro A. Preparation of Few-Layer Graphene Dispersions from Hydrothermally Expanded Graphite. Applied Sciences. 2019; 9(12):2539. https://doi.org/10.3390/app9122539

Chicago/Turabian Style

Vacacela Gomez, Cristian, Talia Tene, Marco Guevara, Gabriela Tubon Usca, Dennys Colcha, Hannibal Brito, Raul Molina, Stefano Bellucci, and Adalgisa Tavolaro. 2019. "Preparation of Few-Layer Graphene Dispersions from Hydrothermally Expanded Graphite" Applied Sciences 9, no. 12: 2539. https://doi.org/10.3390/app9122539

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