A. Rodríguez-Serradeta,
S. Ciftcib,
A. Mikoschb,
A. J. C. Kuehneb,
C. P. de Meloc and
R. Cao-Milán
*a
aLaboratorio de Bioinorgánica, Facultad de Química Universidad de La Habana, La Habana 10400, Cuba. E-mail: robertocao.cuba@gmail.com; Fax: +53 78733502; Tel: +53 78792145
bDWI – Leibniz-Institut für Interaktive Materialien e.V., Forckenbeckstr. 50, D-52056 Aachen, Germany
cDepartamento de Física, Universidade Federal de Pernambuco, 50.670-901 Recife, PE, Brazil
First published on 28th September 2016
Precise assembly of exfoliated graphene sheets on electrode surfaces plays a major role for next generation electrodes to be used in sensors, batteries, fuel cells, supercapacitors and other types of energy storage devices. Here, we achieve vertical assembly of sulfonated, few-layer graphene on a metal surface. The few layer graphene flakes were extensively decorated with dopamine capped iron oxide nanoparticles utilizing electrostatic interactions between the nanoparticles and the graphene sheets. The attached nanoparticles drived the assembly of the graphene nanosheets into a vertical orientation on a gold surface under the effect of a moderate external magnetic field.
The construction of vertically aligned graphene electrodes has been proposed as a possible route to increase the specific surface area, enhancing the charge and power density of a resulting supercapacitor device,5 or the sensitivity of electrochemical sensors.5a
Chemical vapor deposition is commonly applied to grow vertically oriented graphene nanosheets on nickel substrates.6 This process produces very well defined, vertically oriented graphene structures. However, this method is costly, time-consuming and requires high vacuum conditions as well as high temperatures. Such characteristics impede the usage of chemical vapor deposition methods for growing vertically oriented graphene nanostructures over a broad variety of substrates, and thus limiting the construction of more versatile devices.
Alternatively, top-down processing can be applied. It is known that in high magnetic fields reduced graphene sheets, dispersed in a solvent, align parallel to the magnetic field lines. The graphene sheets can then be deposited on a surface, leading to conducting surfaces with very high specific surface area.7 While these are very elegant processes; the lack of susceptible magnetic centers per nanosheet requires the application of high power fields for the alignment, impeding a more widespread application of this method and its rapid transfer into industrial manufacturing processes.7
Hence, a facile wet-processing technology using moderate field magnets to obtain vertically aligned graphene surfaces would enable low-cost and large area processing for next generation batteries, super capacitors, electrochemical sensors and fuel cells. In this sense, we hypothesize that an extensive decoration of graphene sheets with magnetic nanoparticles will enable the use of commercially available permanent magnets to produce vertically oriented graphene surfaces.
Single layer graphene–iron oxide nanoparticles composites are commonly produced by growing the particles on the surface of graphene oxide with the concomitant reduction of the latter.8 In addition; there are a few reports of an alternative procedure for decorating single graphene oxide sheets with previously formed iron oxide nanoparticles (IONPs) by taking advantage of electrostatic interactions.9
Here we propose the decoration of few-layer graphene (FLG) nanosheets with previously formed IONPs utilizing electrostatic interactions to produce a FLG–IONPs composite with a large number of magnetic centers. Under the effect of a magnetic field from commercial neodymium iron boron (NIB) permanent magnets, the IONPs should induce a vertical alignment of the graphene nanosheets, thus facilitating the oriented deposition of FLG on metallic substrates.
We deliberately selected FLG for our study on the magnetic manipulation and orientation. Our material of choice exhibits two important advantages over single layer graphene. First, the chemical modification of the outermost layers does not deteriorate the electronic and conductive properties of the entire sheet, and secondly, the edge-on assembled graphene flakes have sufficient mechanical integrity to stand up perpendicularly to the surface and do not fold back onto the surface, which is often the case when using single layer graphene.10,11
The colloidal stability of the graphene–IONP composite material should play an important role during the magnetically guided vertical assembly onto the electrodes as the necessary rotational motion could be hindered for large aggregates. In this sense, two steps were identified to be crucial for producing a well dispersible FLG–IONP nanocomposite: (i) the preparation of FLG and IONPs precursors (with chemical moieties of opposite charges) both with high colloidal stability, and (ii) a method for mixing both components so as to produce stable dispersion of FLG–IONP nanocomposites.
Succinctly, the objectives of the present study were, first, to synthesize and characterize a sulfonated few-layered graphene (Sulf-FLG) material that could be easily dispersed in polar organic solvents; secondly, to produce super paramagnetic iron oxide nanoparticles capped with a positively charged dopamine protective layer (IONP-Dop); thirdly, to decorate Sulf-FLG with IONP-Dop, producing a Sulf-FLG@IONP-Dop nanocomposite with high colloidal stability; and finally, to vertically assemble the Sulf-FLG@IONP-Dop composite over the surface of gold surfaces and electrodes under the effect of a NIB permanent magnet.
All substrates were created by dividing a silicon wafer coated with a layer of vaporized gold (from Aldrich) into approximately 1 cm2 pieces.
TEM measurements were carried out on a 200 kV FEI Tecnai20.
SEM analyses were carried out on a Hitachi UHR FE-SEM SU9000 and on a Hitachi S-4800 field emission microscope, operating between 2.0 kV and 5.0 kV. Elemental mapping was performed by using an energy-dispersive X-ray (EDX) detector from Oxford (model Xmas 80) attached to the Hitachi UHR FE-SEM SU9000 microscope; operating between 10 kV and 15 kV.
Raman spectra were recorded using a Renishaw Invia Raman microscope equipped with two Peltier-cooled CCD detectors and a Leica microscope with two gratings with 1200 and 1800 lines per mm and band-pass filter optics.
Infrared spectra were recorded using a Rayleigh WQF-510 FT-IR spectrophotometer. Each spectrum was averaged over 32 scans collected at 2 cm−1 resolution. Samples were prepared as of KBr pellets containing approximately 2 wt% of the sample of interest.
Cyclic voltammetric determinations were recorded with a 384B EG&G Princeton Applied Research Polarographic Analyzer Model controlled by home-built software.
Electrochemical measurements of gold substrates were carried out by means of a Teflon cell with a circular hole in the bottom where the gold substrates were in contact with the electrolyte solution. Pictures and details of the construction of such special electrochemical cell can be found in the ESI.† Ag/AgCl (sat) was used as reference and a 2 cm2 platinum sheet linked to a platinum wire was used as counter electrode. Cyclic voltammetry experiments were recorded in a K3[Fe(CN)6] solution (0.01 M) containing Na2SO4 (0.1 M), at a scan rate of 0.1 V s−1.
:
H2SO4 1
:
1) for 12 h and rinsed with bidistilled water. Subsequently, the gold substrates were immersed and stored in HCl (1 M) for 15 min for eliminating any possible oxide from the surface. After the immersion in HCl solution a gold substrate was copiously rinsed with bidistilled water and introduced in 20 mL of a cysteamine methanolic solution (10−2 M) for 24 hours in a closed container. The modified Au substrate was copiously rinsed with bidistilled water before used to assemble the Sulf-FLG@IONP-Dop composite.
The gold substrates modified with a monolayer of cysteamine were introduced in an inclined plastic centrifuge tube containing 10 mL of pure methanol. A commercial permanent magnet (NIB sintered 1.2 T) was located adjacent to the bottom of the tube in parallel disposition with respect to the plane of the gold substrate (see pictures of the system employed on ESI†).
Once the above described system was set up, 100 μL of the Sulf-FLG@IONP-Dop DMSO dispersion were added drop-wise to the methanol containing tube. Under such conditions the Sulf-FLG@IONP-Dop material was deposited on top of the gold substrate within an hour. Subsequently, the methanol inside the tube was removed until the top of the square gold substrate touched the liquid–air interface. 1-Ethyl-3-(3-dimethylaminopropyl)carbodiimide hydrochloride (EDC, 50 mg) and N-hydroxosuccinimide (100 mg) were then carefully added to the solution and the container was opened to allow further evaporation of methanol. When the level of solvent decreased below the bottom of the substrate, the magnet was removed. The substrate was then introduced into an open vial containing methanol (15 mL) and submitted to gentle manual shaking to remove all molecular contaminants. Methanol was then discarded and the previous procedure was repeated two more times. The substrate was finally dried at room temperature in vacuum.
The oxidation of methanol was carried out by immersing each electrode in a methanol aqueous solution (1 M), containing H2SO4 (0.5 M), and at scan rate of 50 mV s−1.
![]() | ||
| Fig. 1 (a) Scheme of synthesis of Sulf-FLG. (b) Transmission electron micrographs of a dispersion of Sulf-FLG. Inset, picture of Sulf-FLG dispersions. (c) Raman spectrum of Sulf-FLG. | ||
The benzenesulfonic acid groups confer solubility to the FLG material in aqueous NaOH solution as well as in DMSO solutions containing tetrabutylammonium hydroxide (see inset of Fig. 1b). These dispersions remained stable over months. TEM analysis reveals Sulf-FLG nanosheets with diameters of 151.30 ± 35.23 nm (Fig. 1b, ESI 1†).
In order to obtain information about the final state of Sulf-FLG material a Raman spectroscopic exploration was carried out (Fig. 1c). The 2D band (commonly denoted as G′) appeared at frequencies above 2700 cm−1, with lower intensity than the G band. This intensity and frequency profile indicates the presence of several graphene layers per FLG platelet.14 The D band at 1350 cm−1 indicates the presence of sp3 carbons, which can be attributed to the incorporation of sulfonate moieties and some oxygen containing defects (see FTIR spectra on ESI 1†). The ratio between the G and D bands clearly indicates a majority of sp2 carbons.14 This observation strongly suggests that only the exposed outermost graphene layers of each FLG platelet turned out to be modified during the process of sulfonic acid functionalization. In a previous report, where a single layer graphene material was submitted to a similar sulfonation process a great amount of sp3 carbons was generated with the consequent deterioration of the electronic properties.15
The IONPs obtained through the solvothermal decomposition of the iron oleate complex resulted of 9.02 ± 1.15 nm as it can be seen in ESI 2.† Such particles resulted capped with a hydrophobic oleylamine monolayer that does not form stable interactions with our Sulf-FLG. In order to increase the affinity of such IONPs towards our Sulf-FLG, the IONPs were submitted to a ligand exchange reaction in DMSO with dopamine hydrochloride. After the interaction with dopamine the IONPs displayed a complex set of IR signals corresponding to phenyl deformation, NH vibrations, and so forth (see ESI 2†). Such result clearly indicates the incorporation of dopamine to the surface of the particles.
The X-ray diffraction pattern and magnetic susceptibility results of our IONP-Dop are consistent with previous analyses of superparamagnetic Fe3O4 particles (see ESI 2†).
Probably, during mixing, highly concentrated IONP-Dop dispersion quickly saturates the negatively charged surface of Sulf-FLG (Fig. 2a). By contrast, when using low concentration of the positively charged particles the formation of insoluble precipitates can result from the association of IONP-Dop acting as bridge between flakes of Sulf-FLG. Regardless of the precise process of mixing, any possible insoluble precipitates formed was discarded and only the suspended Sulf-FLG@IONP-Dop nanocomposites were used in further procedures.
By TEM analysis, it can be observed that our method produced an extensive decoration of Sulf-FLG with IONP-Dop (Fig. 2b). In addition, it can be observed that all IONP-Dop particles are attached to a Sulf-FLG sheet. The large association of IONP-Dop to the outermost layers of Sulf-FLG, produced a large mass increase of each flake of the obtained Sulf-FLG@IONP-Dop in comparison to its IONP-Dop precursor. This difference in weight resulted crucial for rapid separation of Sulf-FLG@IONP-Dop from the excess of IONP-Dop by simple centrifugation at 4000 rpm.
The DMSO dispersions of the obtained Sulf-FLG@IONP-Dop remained stable for weeks.
To conduct the magnetically guided assembly of Sulf-FLG@IONP-Dop onto the gold surface, the simple setup described in Section 2.6 was utilized. The gold substrate (modified with a SAM of cysteamine) was immersed in methanol to reduce the viscosity of the mixture and thus allow the fast magnetic migration of Sulf-FLG@IONP-Dop to the substrate. The substrate was fixed vertically to prevent the Sulf-FLG@IONP-Dop adhesion through sedimentation. The Sulf-FLG@IONP-Dop dispersion in DMSO was added drop wise to the tube, as depicted in Fig. 3a (a photograph of the experimental setup can be found in ESI-3†).
During the assembly, two processes are concurrently occurring: (1) the IONPs attached to each graphene nanosheet orient the FLG along the magnetic field lines, leading graphene flakes perpendicularly aligned with respect to the surface; and (2) Sulf-FLG@IONP-Dop sheets are drawn towards the gold surface, due to the magnetic field exerted by the permanent magnet, leading to the covalent attachment of vertically aligned FLG on the surface (Fig. 3a).
To analyze this assembly, high resolution SEM and EDX elemental analyses were performed (Fig. 3b–e; also, more images can be found in ESI 4†). Vertically oriented graphene nanosheets, with diameters of several hundred nanometers can be observed in the SEM images (Fig. 3b). When a single vertically oriented nanosheet was zoomed-in, its multilayer nature became observable with clearly defined individual edges from each of the stacked graphene sheets (see Fig. 3c). An EDX elemental analysis corroborates that the structures observed in Fig. 3b and c indeed correspond to vertically oriented nanosheets, with signals for C, Fe and O (see Fig. 3d and e). The layer of gold and the underlying silicon substrate are also clearly identifiable (see Fig. 3e, ESI 4†).
A cyclic voltammetry of a [Fe(CN)6]3− solution recorded by means of Au_v_Sulf-FLG@IONP-Dop electrode displayed a 117% increase in current (for the corresponding redox events) when compared to the Au electrodes (Fig. 4a). This result clearly indicates the larger electroactive area of the obtained graphene-based electrode due to the vertical disposition of FLG sheets and its conductive nature (since it is mainly composed by sp2 carbon atoms). Additionally, the electrochemical response of [Fe(CN)6]3− recorded with our Au_v_Sulf-FLG@IONP-Dop electrode displayed a reduction of the peak-to-peak difference (Fig. 4a). Such peak-to-peak reduction has been described to occur when the electron transfer processes take place between [Fe(CN)6]3− and graphene structures through their edges, where the higher density of states near the Fermi level is located.17
The electrochemical properties displayed by Au_v_Sulf-FLG@IONP-Dop demonstrated the suitability of our method for producing new electrodes for electrochemical applications. In this regard, a simple exploratory experiment was made using our Au_v_Sulf-FLG@IONP-Dop electrode for constructing an anode for the oxidation of methanol. For this, such electrode was previously set to electrodeposit Pt (Au_v_Sulf-FLG@IONP-Dop_Pt). For comparison a gold electrode, previously submitted to the electro-deposition of platinum (Au_Pt), was also used as anode for oxidation of methanol.18 In Fig. 4b the voltammograms of the electrooxidation of methanol (1 M) using the Au_v_Sulf-FLG@IONP-Dop_Pt and Au_Pt electrodes are presented. It can be observed that our graphene based electrode displayed a 3.5 fold increment of the current collected during the electroxidation of methanol respect to the Au_Pt one. Such result is also the consequence of the greater electroactive area of the vertically oriented graphene based electrode respect the flat gold electrode.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra15502b |
| This journal is © The Royal Society of Chemistry 2016 |