Vincenzina Barberaa,
Alessandro Portaa,
Luigi Brambillaa,
Silvia Guerraa,
Andrea Serafinia,
Antonio Marco Valerioa,
Alessandra Vitaleb and
Maurizio Galimberti*a
aPolitecnico di Milano, Department of Chemistry, Materials and Chemical Engineering “G. Natta”, Via Mancinelli 7, 20131 Milano, Italy. E-mail: maurizio.galimberti@polimi.it
bPolitecnico di Torino, Department of Applied Science and Technology DISAT, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
First published on 26th August 2016
Electrically conductive flexible carbon papers were prepared, based on hydroxyl functionalized few layer graphene (G-OH). G-OH was obtained from the reaction of KOH with nanosized graphite with a very high surface area (HSAG), greater than 300 m2 g−1, and with high shape anisotropy, with the help of mechanical and thermal energy. Wide angle X-ray and Raman analyses showed that the core of G-OH had the structure of infinite and ideal graphene layers and that the interlayer distance between the few stacked graphene layers was the same as in pristine HSAG. Hydroxyl groups were thus essentially located in peripheral positions. High resolution transmission electron microscopy revealed the presence of graphene aggregates made by a number of layers as low as 6. Stable water suspensions were obtained, with concentrations up to 4 mg mL−1. Mild centrifugation of such suspensions had interesting efficiency as a method for producing few layer graphene: about 35% by mass of pristine HSAG was isolated as few layer graphene from supernatant suspensions. Flexible and electrically conductive carbon papers were prepared by coating a paper support with a G-OH water suspension. This work demonstrates that carbon papers can be prepared without adopting the traditional oxidation–reduction procedure, avoiding harsh reaction conditions, dangerous and toxic reagents, solvents and catalysts.
In light of these facts, various preparation methods have been proposed, after the original micromechanical cleavage.4 Both top-down and bottom-up approaches have been followed. Among the latter ones, preferred appears the epitaxial growth through chemical vapor deposition, mainly on silicon carbide or metals.9–12 Top-down methods could appear promising in view of their industrial development, as traditional graphite could be in principle used as starting materials. Typical top-down method involves the oxidation of graphite or graphitic nano-fillers, made by few layers of graphene, to graphene oxide (GO)13–17 and the successive reduction. Precise structure of GO is still unknown.18 It was reported19 that carbonyl and carboxyl groups are on edges of basal planes and hydroxyl and epoxide groups are on the surface. Such functional groups are able to promote the dispersion of the modified graphitic layers in water and other polar solvents, allowing the preparation of coating layers. This is reported in the literature as a key step for the preparation of graphene papers, one of the investigated applications of graphene. However, this approach to the preparation of graphene papers is affected by some main drawbacks. Oxidation, typically performed through Staudenmaier13 or Hummers14 methods, requires explosive oxidation agents, strong acids, harsh reaction conditions and long reaction times. Moreover, it is widely acknowledged that oxidation leads to extensive disruption of sp2 hybridization of graphene layers, with loss of properties. The reduction step is thus ineluctable, to restore the graphene layers and in turns, to allow the preparation of conductive graphene papers. However, chemicals for the reduction could be toxic, hazardous and expensive, such as, for instance, those based on hydrazine.20–22 Moreover, high temperature is needed for thermal reduction.23,24 Other reduction methods have been recently reported, for the obtainment of graphene papers25–29 Thermal treatments at 700 °C have been applied to graphene oxide paper, under argon or hydrogen atmosphere.25 Supercritical ethanol was used.26 Ar+ ion irradiation was performed on GO paper obtained by vacuum drying of water dispersion on stainless steel plate.27 Reduced graphene-oxide (r-GO) papers, prepared by vacuum filtration, were treated with Ar/H2 atmosphere.28 GO film on a PET substrate (PET) was reduced with a magnetron sputtered Sn metal layer.29 However, all the approaches documented in the literature for the preparation of graphene papers are based on two steps: preparation of GO and successive reduction.
In the light of this situation, it appears worthwhile to make research for an efficient, simple method for the preparation of graphene and/or few layers graphene, suitable for the preparation of flexible and conductive carbon papers. Ideally, such method should be based on a single step, in line with the basic principles of green chemistry. The introduction of functional groups on graphene layers, preferentially oxygenated functional groups, would be highly desirable, as well as the preservation of the sp2 nature of carbon atoms.
This work had these objectives. Oxidation of a nanosized high surface area graphite (HSAG) was pursued through the reaction with KOH, with the aim to selectively introduce hydroxyl groups on edges of graphene layers, leaving substantially unaltered the bulk crystalline order and thus the aromatic nature of the layers. Further aim was to promote exfoliation of HSAG, achieving the preparation of few layers graphene. Elimination of hydroxyl groups was not pursued. On the contrary, functionalized graphene layers were used as macro-substrates for further functionalization reactions. The one with isocyanate to form urethanes is reported as an example. Further objective was the obtainment of few layers graphene suspensions in ecofriendly solvents such as water and the preparation of flexible and electrically conductive carbon papers by deposing G-OH suspensions on different substrates.
To pursue such objectives, strategic was the selection of the nanosized graphite. HSAG had a very high surface area, higher than 300 m2 g−1, and a high shape anisotropy, defined as the ratio between the crystallites size in directions perpendicular and orthogonal to the layers.30 In particular, HSAG had a low number of stacked layers, about three tenths. Reaction of HSAG with KOH was performed in a single step, with the help of either mechanical or thermal energy. Few layers graphenes with OH functional groups (G-OH) were characterized by means of elemental and thermogravimetric analysis (TGA), high resolution X-ray photoelectron spectroscopy (XPS), infrared (IR) and Raman spectroscopy, wide angle X-ray diffraction (WAXD), high-resolution transmission electron microscopy (HRTEM) and cyclic voltammetry. Dispersions in water of G-OH were prepared and investigated through dynamic light scattering (DLS) and UV spectroscopy. Mild centrifugation of such dispersions and recovery from supernatant was studied as a method to obtain few layer graphene. Thin films were prepared on paper substrate starting from a water suspensions of G-OH. Flexibility and electrical conductivity of the obtained carbon papers were investigated.
C bend), 856 (out of plane,
C–H bend, monosubst) cm−1.
C bend), 850 (out of plane,
C–H bend, monosubst) cm−1.
C bend), 850 (out of plane,
C–H bend, monosubst) cm−1.
C bend), 721 (out of plane,
C–H bend, monosubst) cm−1.
Dhkl = Kλ/(βhkl cos θhkl)
| (1) |
The yield of few-layer graphene (FLG) was determined by analysing the weight percentage of all of exfoliated flakes after separation by centrifugation.
![]() | ||
| Fig. 1 Block diagram for the preparation of G-OH adducts: by either thermal (G-OH-T) or mechanical energy (G-OH-M). | ||
Preparation of G-OH via mechanical treatment was carried out adopting two KOH/HSAG ratios: about 20
:
1 and 10
:
1 (method A and method B, respectively, in the experimental part). In this paper, samples obtained with method B are discussed. G-OH samples were washed until neutral pH of washing water. Characterization of three representative samples, from thermal, mechanical and thermal-mechanical treatments, is discussed as follows.
Chemical composition of the samples was investigated by means of elemental analysis, TGA, XPS and FTIR analysis.
Results from elemental and thermogravimetric analysis and thermographs of HSAG and G-OH-M are reported as ESI.† Data from elemental analysis are in Table S.1.† Pristine HSAG was found to have the following chemical composition (% by mass): carbon 96.8, oxygen 4.2. The oxygen content of G-OH adducts, as expected, increased to about 5.0 and 6.1, for G-OH-T and G-OH-M, respectively. Largest amount of oxygen from 9 to 16% was introduced by adopting larger KOH/HSAG ratio (method A) for the mechanical treatment. Method B, based on lower KOH amount, was used in order to modify pristine HSAG only to a minor extent.
TGA revealed lower mass loss for HSAG, below 700 °C, than G-OH samples, as it is shown by data in Table S.2 and by comparison of traces in Fig. S.1 (HSAG) and S.2–S.4† (G-OH). For G-OH samples, a three-step decomposition profile can be observed. The first mass loss below 150 °C can be attributed to the water removal. The second and third mass losses, that are due to decomposition of oxygen-containing groups and removal of alkenylic groups, occur in the temperature range between 150 °C and 700 °C. Combustion with oxygen of graphite occurs for temperatures higher than 700 °C. The low mass loss of HSAG below 700 °C indicates the presence of a minor amount of sorbed water and of functional groups that can undergo degradation. For G-OH samples, the larger amount of mass loss at T > 150 °C is accompanied by larger mass loss at T < 150 °C. It could be hypothesized that G-OH containing larger amount of oxygenated functional groups holds larger amount of water. It is finally worth observing that the onset temperature of the last degradation step becomes appreciably lower in G-OH samples than in HSAG sample, presumably as a consequence of combustion of edge carbons with oxygen-containing functional groups.
Type and amount of atoms and functional groups in pristine HSAG and in G-OH were then studied through wide scan XPS spectra (Fig. 2a). As expected, in both spectra the main signals are due to C 1s and O 1s (binding energy B.E. 285 eV and 532 eV respectively). After treatment with KOH the modified graphite shows also two small noisy signal at B.E. around 400 eV and 385 eV, attributable to N 1s and K 2s. The O 1s/C 1s atomic ratio of the pristine graphite is 0.04 (slightly higher than the 0.03 value reported in the literature for graphite), corresponding to an atomic O percentage of 4.2%. Treatment of KOH affects the oxygen content: in G-OH its atomic percentage is in fact 6.4%, and the O 1s/C 1s ratio increases to 0.07.
Narrow scan spectra for each element were recorded, C 1s and O 1s envelopes were evidenced and different contributions were identified by a curve fitting processing. As shown in Fig. 2b, C 1s (main peak at B.E. = 284.8 eV) of both samples have a broad, asymmetric tail towards higher binding energy and one satellite signal, few eV from the main peak, related to the π–π* transition (shake-up at around 290.9 eV): these two features are found when a high concentration of sp2 carbon is present. In the envelopes of C 1s, the peak centered at 284.8 eV is related to C sp2; the other peaks at 285.3 eV and 287.3 eV are likely to characterize C–O and C
O functions and their intensity is changed by the KOH treatment, although one can not exclude the presence of C–H environments and other defects. The relative portion of these functions, i.e. the C–O/C
O ratio, changes from 2.2 to 5.4, which means that the amount of C–O groups increases with the functionalization reaction. These data are confirmed by the analysis of the O 1s envelope: on both samples, in fact, two main peaks are present, corresponding to C–O type oxygen at 533.1 eV (hydroxylic groups, non-carbonyl – ether type – oxygen atoms in esters and anhydrides) and to C
O type oxygen at 531.1 eV (carbonylic oxygen in esters and anhydrides).33
Functional groups present in HSAG, G-OH and GO samples were investigated with IR measurements on samples prepared in a Diamond Anvil Cell (DAC). In Fig. 3 the IR spectra of HSAG (a), G-OH-T (b), G-OH-TM (c), G-OH-M (d) and GO from Staudenmaier oxidation (e) are reported.
All the spectra in Fig. 3 are characterized by an increasing background toward high wavenumbers due to diffusion/reflection phenomena of the IR light by the particles of the sample. The spectrum of HSAG (Fig. 3a) is characterized by the band at 1590 cm−1 which is the peak of graphite and graphene materials, assigned to E1u IR active mode of collective C
C stretching vibration. In the spectra of G-OH samples (Fig. 3b–d) the C
C peak is still observed along with new bands. These new bands can be assigned to the absorption of different –OH groups bonded to the graphene sheets. The corresponding peaks are located at: (i) 3400 cm−1, assigned to the –OH stretching vibrations of hydrogen bonded hydroxyl groups; (ii) 1388 cm−1, assigned to the out of plane vibration –OH groups; (iii) 1110 cm−1, assigned to C–O stretching vibration and (iv) 970 cm−1, assigned to the in plane phenyl–O–H bending.
The spectrum of GO (Fig. 3e) shows peculiar features different from G-OH samples. The band at 3400 cm−1 (–OH stretching) is broad and stronger. Carboxylic groups originates the large and strong band located at 1727 cm−1. The C
C peak becomes very strong due to the polarizing effect of the –C
O groups bonded to the graphene backbone. Other strong and broad features occurring at 1250 cm−1 and 877 cm−1 can be assigned to the ring breathing and the C–O–C bond stretching of epoxide groups respectively, together with other vibrations of C–O–C and –OH functionalities. The peak at 1044 cm−1 could be ascribed to the stretching vibration of S
O bonds. It is worth noticing that the IR spectrum of GO from Staudenmaier method (Fig. 3e) shows the presence of oxygenated functional groups, mainly carboxyl and epoxide groups, that are not present on G-OH samples.
Structure of HSAG and G-OH were studied by means of WAXD, Raman and TEM analyses.
Fig. 4 shows WAXD patterns taken on powders of HSAG (Fig. 4a), G-OH-T (Fig. 4b), G-OH-M (Fig. 4c) and G-OH-TM (Fig. 4d). Moreover, it is shown the pattern of a GO sample obtained through the Staudenmaier method (Fig. 4e).
In pristine HSAG, crystalline order in the direction orthogonal to structural layers is revealed by two (00l) reflections: 002 at 26.6°, that corresponds to an interlayer distance of 0.338 nm and 004 at 54.3°. Such interlayer distance is slightly larger than the one of ordered graphite samples (d002 = 0.335 nm).34 The in plane order is shown by 100 and 110 reflections, at 42.5° and 77.6° respectively. By applying the Scherrer equation (see the Experimental section) to (002) and (110) reflections, respectively, the out of plane (D⊥) and the in plane (D∥) correlation lengths were calculated. Values were 9.8 nm for (D⊥) and 30.2 nm for (D∥). The in-plane correlation length is thus larger than the out-of-plane correlation length. From the values of (D⊥) and of the interlayer distance, the number of stacked layers was estimated to be about 35. These results reveal the turbostratic nature of HSAG,30 that has however remarkable crystalline order inside the structural layers. HSAG appears thus a good graphite grade for the preparation of conductive coating layers, provided that the sp2 nature of carbon atoms is not perturbed.
(002) reflection in the patterns of G-OH samples remains at the same 2θ value, indicating that the oxidation reaction did not promote expansion of the interlayer distance. There is, instead, a large change in the position of (00l) reflections in the pattern of GO from oxidation with the Staudenmaier method. (001) reflection is at 2θ = 10.45 and corresponds to an average interlayer spacing of 0.9 nm. These differences are justified by the presence of oxygen-containing functional groups attached on both sides of the graphene sheet. The number of stacked layers in G-OH samples was calculated by applying the Scherrer equation to 002 reflection and was found to be between 20 and 25. 100 and 110 reflections, due to the in plane order, are clearly visible in G-OH patterns. From these findings, it can be commented that the reaction with KOH promotes the partial exfoliation of HSAG without altering the in plane order.
Raman spectroscopy is widely employed for the study of carbonaceous materials.35–41 In Raman spectra, two peaks, named D and G and located at 1350 cm−1 and 1590 cm−1 respectively, are in particular investigated. The G peak is due to bulk crystalline graphite (graphene), whereas the D peak appears in the presence of either structural defects, such as holes, sp3 or sp carbon atoms, free radicals, distortions from planarity, grafted functional groups or confinement (e.g. by edges) of the graphitic layers.38–41 It is worth reminding that graphitic layers have finite dimensions and irregular boundaries with, for example, dangling bonds. Graphitic powders that experienced ball milling, as HSAG, revealed in the Raman spectrum evident D line, whose intensity increased with the grinding time.41 In this work, Raman spectroscopy was applied to the structural characterization of HSAG and G-OH-M, and the spectra are shown in Fig. 5a and in Fig. 5b, respectively. Analysis was performed on a good number of spots and spectra can be considered representative of the investigated samples.
G and D bands are present in both spectra, with similar intensity. The high intensity of the D band can be interpreted taking into account what reported in the literature for Raman spectra of ball milled graphite.42 The electronic and vibrational properties of a core of graphitic samples sufficiently far from the edge can not be distinguished from those of infinite and ideal graphene layers. A crown confined region, close to the edge, is electronically perturbed and the electronic and vibrational structures are different from those of the bulk materials. The pronounced D band is thus due to the different types of molecular disorder. However, there are not indications that the reaction of HSAG with KOH appreciably alter the structure of the graphitic layers.
The structure of G-OH adduct was investigated also by performing HRTEM analysis on samples isolated from water supernatant suspensions (see below in the text for the discussion on water suspension), after centrifugation for 10 min at 2000 rpm and for 15 min at 9000 rpm. Fig. 6 shows HR-TEM micrographs at lower and higher magnifications.
Micrographs at lower magnification in Fig. 6a and c reveal that the lateral size of G-OH adducts is of the same order of magnitude in samples isolated after centrifugations at different rpm and for different times. Hence, there is not indication that the milling step cause appreciable breaking of the graphitic layers. Micrographs at higher magnification allow to visualize the stacks of graphene layers that are disposed with a lateral side perpendicular to the beam. Thanks to this disposition of the nano-stacks, it is possible to estimate the number of stacked graphene layers. Fig. 6b shows stacks isolated after centrifugation for 10 min at 2000 rpm: are visible stacks (indicated in the boxes) of about 1.7–4.8 nm, with 6 to 15 stacked graphene layers. As revealed by the exam of a good number of micrographs, such stacks are the most abundant ones in a stacks' population that contains either little larger of little lower number of layers. Fig. 6d shows a stack (indicated in the box) isolated after centrifugation for 15 min at 9000 rpm: is of about 1.7 nm and contains about 6 stacked graphene layers. Such a stack was very frequently observed in the analyzed micrographs after centrifugation for 15 minutes.
Reaction of high surface area nanosized graphite with KOH allows thus to introduce an appreciable amount of hydroxyl functional groups (from 4 to 6% by mass), promoting partial exfoliation but maintaining the same interlayer distance between the graphene layers, without altering the bulk crystalline order of the graphitic layers. It appears thus a reasonable hypothesis that hydroxyl groups are located on the edge carbon atoms. A mechanistic pathway can be hypothesized for this reaction. The insertion of hydroxyl anion can occur on peripheral carbon and the following two step mechanism can be hypothesized: (i) a nucleophilic initial addition of an hydroxide ion to the peripheral aromatic ring followed by (ii) re-aromatization processes. In the frame of this mechanism, a negatively charged intermediate would be formed. Such intermediate would exist in a considerable number of resonance hybrids. Such large stabilization could be considered a sort of “wave effect”. Driving force would be the aromatization of the graphene layer (Fig. S.5†).
Dispersions of G-OH in water were prepared through a mild sonication. Results obtained with G-OH-TM are reported as follows. Dispersions were with different concentrations (0.01, 0.02, 0.05, 0.1, 0.2, 0.3, 0.5, 1.0 and 4.0 mg mL−1). Fig. 7 shows dispersions in the range from 0.01 to 1 mg mL−1.
Fig. 8 reports results from UV-Vis absorption analysis. Fig. 8A shows that the absorbance monotonously increases with G-OH-TM concentration and Fig. S.6† demonstrates a linear correlation between absorption and G-OH-TM concentration, thus revealing that water suspension of G-OH-TM were prepared. Fig. 8B shows that the absorbance detected for the freshly prepared suspension (1 mg mL−1) did not change after 1 week storage. Traces in Fig. 8C indicate that the centrifugation at 2000 rpm for 10 minutes led to a slight decrease of absorbance.
Stability of G-OH-TM suspension was verified also at a concentration of 4 mg mL−1, after storage for 4 weeks.
Average size of G-OH-M and G-OH-T aggregates in water suspensions was investigated by means of dynamic light scattering (as reported in the experimental part). Table 1 shows the average size of particles, for water suspensions of HSAG and of both the adducts.
Introduction of oxygenated functional groups leads to the reduction of HSAG aggregates size. G-OH particles in the supernatant suspensions (both from mechanical and thermal treatments) appear to have lower average size than particles of HSAG and GO. In particular, the mechanical treatment appears to promote the formation of smaller G-OH aggregates. It could be also commented that the presence of large amount of functional groups can favour the aggregation of GO particles.
Preparation of G-OH water suspension was investigated as a method for producing few layer graphene. As described in the experimental part, 1 mg mL−1 G-OH suspension was centrifuged at 2000 rpm for 10 minutes, obtaining a supernatant suspension containing stacks of graphene with layer number ranging from 6 to 15, as observed via HRTEM and confirmed by WAXD analysis. The amount of few layer graphene in the supernatant suspension was about 35–40%.
As mentioned in the introduction, objective of this work was the preparation of flexible and electrically conductive carbon papers by deposing G-OH suspensions on different substrates, avoiding the reduction step. G-OH layer was deposed on paper using a bar coater (as described in the experimental part, with 6, 12, 24, 40
100 μm bars), obtaining flexible carbon paper. Two types of suspensions were mainly investigated: 1 mg mL−1 suspension as prepared by sonication and suspension after centrifugation at 2000 rpm for 10 minutes. Structure of the graphitic coating on paper was studied by means of WAXD analysis. Coating layers obtained with both types of suspensions revealed a number of graphene layers (evaluated from peak shape analysis, as reported above) close or higher than that detected on pristine G-OH. In Fig. S.7† are reported patterns of G-OH-M (a), paper support (b) and graphitic coating layer (c). D002 correlation length of the graphitic layer of about 17 nm (corresponding to about 50 stacked layers) was calculated. This result could be expected, as the coating procedure favors the stacking of graphene layers and indicates that the key aspect for the preparation of few layers graphene coating is the selection of the starting suspension: supernatant suspension after centrifugation should be used.
Electrical conductivity of G-OH coated paper was measured with the four point probe, as described in the experimental part, and data are shown in Table 2.
Very high resistivity was measured for the GO layer from Staudenmaier oxidation, as expected. In fact, only the reduction step is able to restore, though if not completely, the sp2 hybridization of the carbon atoms. Conductive HSAG layer was not obtained, probably as a consequence of the large amount of binder (carboxymethylcellulose) needed to disperse HSAG. The electrical conductivity detected in the G-OH layer confirms that carbon atoms have largely prevailing sp2 hybridization, as HSAG functionalization occurred on peripheral positions. Defects on the edges of graphene layers in pristine HSAG and hydroxyl functionalization can be considered to adversely affect the electrical conductivity of carbon papers.
Electrical properties of G-OH were also investigated by cyclic voltammetry. In the literature, CV was used to show the good electroactivity of G-OH (from ball milling with KOH).32 Peak separation, measured for the [Fe(CN)6]4/[Fe(CN)6]3 redox pair, was found to significantly decrease when the glassy carbon electrode was coated with G-OH and this was taken as indication of significant increase of electron transfer rate. This CV experiment and the above reported four point measurements show that G-OH with substantially unaltered graphitic structure is electrically active. In this work, CV measurements were taken (graphs are in Fig. S.8†) on 0.1 M Et4NBF4/10 mM K4Fe(CN)6 solution (a), obtaining the expected cyclic voltammogram, and on a 0.125 mM G-OH suspension/0.1 M Et4NBF4 (b), without observing any detectable signal, as it should be expected for a graphite with very minor amount of oxygenated functional groups. On the contrary, cyclic voltammograms are reported in the literature for graphite oxides obtained through standard oxidation methods,43 that means GO with remarkable amount of functional groups, such as GO from Staudenmeier oxidation. These findings are in line with the other results of this work: OH functionalization allows the preparation of water graphite suspensions and of carbon papers, without affecting graphitic electrical properties. CV was also carried out on the 0.1 M Et4NBF4/10 mM K4Fe(CN)6 solution, in the presence of 0.125 mM G-OH (Fig. S.8c†). Shift of [Fe(CN)6]4/[Fe(CN)6]3 CV curves suggests that G-OH can be used to modify the behavior of redox pairs.
Results reported in this manuscript appear to pave the way for easier applications of graphene and few layers graphene, through the selective reactivity of aromatic rings in the graphene layers.
Origin of Stability and Potential for Magnetism in Carbon Materials, J. Am. Chem. Soc., 2005, 127, 5517 CrossRef PubMed Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra19078b |
| This journal is © The Royal Society of Chemistry 2016 |