Xiao
You‡
*a,
Sangita
Maharjan
b,
Kizhanipuram
Vinodgopal
b and
Joanna M.
Atkin
*c
aDepartment of Applied Physical Science, University of North Carolina at Chapel Hill, Chapel Hill, USA. E-mail: youxiao@westlake.edu.cn
bDepartment of Chemistry, North Carolina Central University, Durham, USA
cDepartment of Chemistry, University of North Carolina at Chapel Hill, Chapel Hill, USA. E-mail: jatkin@email.unc.edu
First published on 3rd January 2024
Graphene oxide (GO) has attracted substantial interest for its tunable properties and as a possible intermediate for the bulk manufacture of graphene. GO and its reduced derivatives display electronic and optical properties that depend strongly on their chemical structure, and with proper functionalization, GO can have a desirable bandgap for semiconductor applications. However, its chemical activity leads to a series of unclear chemical changes under ambient conditions, resulting in changes in color and solubility upon exposure to light. In this paper, we study the properties of fresh and spontaneously reduced GO under ambient conditions using tip-enhanced Raman spectroscopy (TERS) to map its nanometer scale chemical and structural heterogeneity. We observe different types of defect sites on reduced GO (rGO) by spatially mapping the D to G band peak ratio and D and G band spectral positions. The higher spatial resolution and out-of-plane polarization compared to conventional micro-Raman spectroscopy enables us to resolve unusual features, including D-band shifting on rGO. Based on statistical analysis of the spatial variations in modes and theoretical calculations for different functional groups, we conclude the reduction mechanism of GO is a self-photocatalytic reduction with the participation of water and visible light, in which the rate determining step is electron transport through the metal substrate and ion diffusion on the GO surface. These results demonstrate that TERS can reveal structural and chemical details elucidating reduction mechanisms, through the examination of samples at different time points.
Micro-Raman spectroscopy, though a powerful tool for the study of sp2 carbon systems due to the large Raman cross-section of delocalized conjugated bonds, is limited to providing an averaged signal within the large laser spot size. This resolution limit has severely hindered characterization of the structural and chemical heterogeneity in GO and rGO, which exhibit variations over length scales ranging from the molecular to tens of nanometers. In Raman spectra of pristine graphene, characteristic peaks include the G-band around 1580 cm−1 and the G′ (or 2D) band at 2700 cm−1, but in chemically or structurally disordered graphene and graphene oxide a D-band emerges close to 1350 cm−1. Researchers often employ the relative intensity of the D and G bands (ID/IG) as an indicator of disorder, particularly for assessing the degree of reduction in GO. Nonetheless, samples with the same level of oxidation can exhibit significantly different ID/IG ratios, as different reduction methods can either reduce or increase the number of defects.18–20
Tip-enhanced Raman scattering (TERS) combines atomic force microscopy (AFM) with a plasmonically-enhanced Raman signal, enabling spatial resolutions from a few nanometers down to single chemical bond levels.21–26 Therefore TERS presents a promising technique for resolving structural disorder in GO systems. The exceptional spatial resolution of TERS has enabled imaging of local defects, strain, and grain boundaries in graphene and carbon nanotubes with exquisite sensitivity.27–31 However, for GO and rGO, TERS faces a bottleneck in data analysis and interpretation due to the high degree of disorder, and modification of the selection rule in TERS compared to conventional Raman spectroscopy. Previous measurements on these systems have relied on samples with defined moieties such as carboxyl-modified GO, and studied the absolute and relative intensities of the D and G bands.32,33 In complex systems, this form of intensity analysis in TERS may not reveal chemical information, due to the sensitivity of the Raman enhancement to variable sample thickness, roughness of the underlying metal substrate, and variations in the coupling of tip and surface.34
Here, we therefore investigate the nanoscale spectral properties of graphene oxide and reduced graphene oxide, with particular focus on frequency shifts in the D and G band, which we find to be related to disordered topographic features. We perform statistical analysis on the spatial variations in the peak positions, and use density functional theory (DFT) to associate the variations with chemical defects and functional groups. Based on the correlated topographic and spectral variations we observe, we propose that the self-reduction mechanism of GO is a photocatalytic process with the participation of water and visible light, which produces a more defected sp2 lattice in folded, wrinkled, or multilayer regions than in single layer regions. This analysis approach demonstrates the potential of TERS for understanding disorder in complex functional materials and their structural evolution, as well as providing a pathway to controlled reduction in graphene oxide systems for catalytic and electronic applications.
For our experiments, we dissolve freshly synthesized GO in ethanol and deposit it onto Au sputtered substrates. The good solubility of GO in ethanol yields a well dispersed sample, with single to few layer regions. One sample was measured shortly after deposition. A second sample, designed rGO1, was stored under ambient conditions for seven days before measurement. The third sample, rGO2, was stored for 30 days to achieve partial reduction. Incomplete reduction leaves behind a moderate amount of residual oxygen-bearing species and defects, which provide an ideal platform for study of GO's chemical properties.
To estimate the chemical composition, we conducted X-ray photoelectron spectroscopy (XPS) on GO and spontaneously reduced GO (rGO), with chemically reduced GO as a standard for comparison (crGO). Previous studies on the C 1s XPS spectrum of GO show four components that correspond to carbon atoms in different functional groups: the non-oxygenated ring C, the C in C–O single bonds, the carbonyl C and the carboxylate carbon (O–CO) at 284.6, 286.2, 287.8 and 289.0 eV respectively, showing in Fig. 1a.1 In crGO, the C 1s peak at around 287 eV decreases in intensity and is red-shifted following chemical reduction, indicating a chemical conversion of the C
O and C–OH groups to a new chemical species. In the rGO sample, however, this peak is significantly weaker, without peak shifting. The decrease of these peaks indicates loss of oxygen and the related functional groups. By curve fitting analysis of the XPS results, we can determine the percentage of the various functional groups (Fig. 1b) and oxygen (black). These results provide proof of partial reduction in rGO, though less so than the chemically-reduced crGO.
Micro-Raman spectroscopy can provide spatially-resolved spectral information, and is therefore a valuable complementary tool to XPS. However, although XPS clearly registers significant variations in chemical composition in samples from different preparations, the Raman spectra all look nominally very similar (Fig. 1c), with only a small decrease in ID/IG ratio for the chemically reduced sample. Therefore this commonly used ratio is not a reliable parameter for the degree of chemical reduction. We overcome this limitation of far-field optical probes by employing tip-enhanced Raman scattering (TERS), using a commercial atomic force microscope (AFM)-Raman system (Horiba). We use silicon AFM tips, sputtered with gold, to generate the necessary plasmonic enhancement for high sensitivity and spatial resolution. Details on the tip fabrication and measurement parameters are provided in the ESI.†
![]() | ||
Fig. 3 (a) AFM image of sample rGO2 on gold surface. TERS mapping of the region in (a) based on (b) the G-band position, (c) D-band position, and (d) ID/IG ratio. |
These spatial variations are even more clearly resolved in the sample rGO2, reduced for 30 days. In particular, the flake on the bottom right shows a clear distinction between a folded and flat region in the D-band position, while the D/G ratio remains relatively constant aside from the wrinkled area.
For comparison, TERS images of the fresh GO are shown in the Fig. S3 (ESI†). In the area marked in the AFM image of GO (Fig. 3a), D and G band peak positions and their intensity ratios (ID/IG) are shown in Fig. 3b–d. While the ID/IG ratio map highlights the edge and wrinkles on the GO flake, similar to the features of the rGO, the band position map (Fig. 3b and c) is more homogeneously distributed than rGO, with no regions of clearly distinct spectral variations. In spite of this lack of consistent peak shifts, the spectral distributions still suggest significant chemical heterogeneity. We note that the intensity maps for the D- and G-band, shown in Fig. S4 and S5 (ESI†) for GO and rGO respectively, are strongly correlated. That is, the regions of high intensity for D-band occur at the same spatial positions as for the G-band, suggesting that the only distinguishable intensity variations are due to differences in tip-sample coupling, likely to be dominated by the roughness of the substrate.
We statistically analyzed the spatially-resolved spectra to collect peak position data in the range 1250–1400 and 1550–1650 cm−1 from the TERS maps. To better resolve the detailed chemical structures of reduced GO, we separated the observed D-band positions from different regions of the rGO1 (Fig. 2c) and rGO2 samples (Fig. 3c) and the GO flake (Fig. S3d, ESI†). The plotted histograms of D-band and G-band peak positions in these two samples are compared in the supplement. We observe that the D-band has a large peak shift in rGO2 (red) compared to GO (black). The D-band peak frequencies are also significantly more distributed in rGO1 and rGO2 than in GO, implying increased chemical heterogeneity. The G peak position in contrast is only slightly red-shifted in rGO2, and displays a similar distribution. Since G band peak positions are related to properties such as the number of graphene layers and point defects and have been previously discussed,35,36 we focus here on the D-band.
Based on the box plots in Fig. 4, we can see that the spread in peak position is relatively narrow in GO, and shows no systematic variation between flat, edge, and wrinkled regions. In rGO, in contrast, the different regions show very different average peak positions and distributions. The spread in the peak position even for a flat region in rGO1 and rGO2 is considerably larger than a comparable region in GO, although the averages are similar. In rGO1, the edge region has a blue-shift in the D-band peak position, with an even larger shift in the folded region. A similar trend is observed in rGO2, but with an even larger shift at the edges and folded regions, with the mean peak position moving from 1350 cm−1 to 1360 cm−1 with increased reduction time. This observation suggests that reduction proceeds inhomogeneously across the surface of GO, occurring preferentially at folded regions. Both edges and wrinkles also have large spreads in peak positions, consistent with the higher density of functional groups expected in these regions compared to flat regions.
Understanding the chemical and structural variations corresponding to the differences we observe experimentally is challenging. Key elements responsible for the emergence of the D band in graphene include lattice defects, phonon–phonon interactions, and double resonance processes that involve electron–phonon interactions and scattering by defects.38–40 Moreover, the double resonance process involves four distinct double resonance Raman scattering mechanisms, comprising outer and inner scattering processes, as well as scenarios where electrons are initially scattered by phonons (referred to as “phonon-first”) or by defects (termed “defect-first”). These processes give rise to two sub-bands within the D band and involve intricate interactions between electrons, phonons, and defects that significantly influence the characteristics of the D band.41 Collectively, these elements define the distinct spectral signature of the D band in Raman spectroscopy. These factors cannot be easily deconvolved in a complex, disordered material such as rGO. To provide further insights into the observed shifts in the D band, we therefore employed density functional theory to model the functional groups present in GO at the atomic and electronic levels.
As a first step, we must establish which defects and moities are likely to be most significant in our samples. X-ray diffraction and scanning tunneling microscopy structural studies of GO have indicated that the center of GO is mainly functionalized with hydroxyl and epoxy groups,10,46 with other groups such as ketones and ethers along the edges.47 Considering that the size of a GO flake is usually in the μm range, the influence of the more diverse edge groups on the overall phonon mode can be neglected in our calculations. Considering only the center of a GO flake, we can idealize GO and rGO as a graphene lattice with defects that contain C–O and CO bonds at a ratio of one oxygen atom per eight carbon atoms, consistent with our XPS results. Thus we build two models for GO, treating it as a hexagonal graphene lattice with one added oxygen-containing group, either epoxy or hydroxyl, every eight carbon atoms (Fig. 5a).
For rGO, electron microscopy studies suggest that it consists of defect-free graphene areas with sizes of a few to ten nanometers interspersed with defect areas with disordered carbon networks (remnants of oxygen containing groups), individual ad-atoms or substitutions, isolated topological defects (single bond rotations or dislocation cores) and holes and edge reconstructions.48 Previous studies also show that the reduction of GO results in both a restored C-sp2 region, and C-sp2 with defects.46 The most common defects in rGO are the C-sp3 and carbon vacancies. Thus the two rGO models we employ consist of graphene lattices with added C-sp3 and removal of one carbon atom every eight carbon atoms, respectively.
Prior to calculation of the theoretical Raman spectra, the most probable geometries of the sample are modeled in Materials Studio (2017 R2). These structures are geometry optimized and then Raman modes are calculated using the CASTEP package with GGA functional and the PBE basis set.
Our calculation of pristine graphene produces the expected E2g mode at a vibrational frequency 1588.4 cm−1, which is in good agreement with the experimental Raman spectrum of pristine graphene (G-band peak position at 1587 cm−1) (Fig. 5b, black.) A complete restoration of the sp2 π-conjugated network in rGO should generate only the G band.
The calculation of Raman spectra of GO with epoxy groups (red) and hydroxyl groups (blue) shows a shift in the G-band and the appearance of additional modes in the standard D-band spectral range of 1300–1400 cm−1. For the rGO models, the disturbance in the sp2 π-conjugated network due to missing carbon atoms causes the G-band to be significantly weaker than pristine graphene, and a strong peak between 1350–1400 cm−1 appear (green). The relative strengths of the D and G band in this calculation suggest that the true concentration of C vacancies is lower than our model. The addition of sp3 defects in the lattice produces two strong modes, at around 1100 and 1300 cm−1 in addition to the G-band (purple). We did not observe a strong vibrational mode at 1100 cm−1 in our samples, suggesting that sp3 defects do not appear in significant quantity. These modes have been previously observed, however, in amorphous and nanocrystalline sp3 carbon,49 and thermally reduced GO.50
Our experimentally-observed D-band likely arises due to contributions from epoxy, hydroxyl, and C vacancies, producing broad, overlapping spectra where even the small probe volume in TERS likely contains many defects. These DFT results can provide hints of the origin of our observed D-band shifting, in spite of our inability to separate all the possible spectral contributions. In particular, we note that the conversion of some epoxy groups to hydroxyl would blue-shift the D-band, while carbon atom vacancies result in greater blue shifts than either of the functional groups.
Dimiev et al. proposed several transformation schemes for GO production during exposure to water, suggesting the formation of double bonds at the expense of tertiary alcohols and C–C bond rupturing.51 While their mechanism explains certain experimental observations, such as the extension of conjugated areas and the acidic properties of GO, our experimental results, including XPS and Raman data (Fig. 1a),indicate a clear trend of the reduction of the oxygen-to-carbon ratio and an increase in sp2 domain and C vacancy defects. This observation, supported by TERS and DFT results, differs from Dimiev et al.'s proposed mechanism. Moreover, both the oxidation of graphite and the reduction of GO resulted in the production of CO2 gas, as confirmed by gas chromatography.51,52 However, the stability of the C–C bond at room temperature suggests its resilience to breakage without additional energy input. This implies that the generation of CO2 in the reduction process depends on the excitation energy and stability of the local structure. Notably, the mechanisms proposed in earlier studies do not provide a satisfactory explanation for the observed release of CO2.
Previous studies have highlighted GO's involvement in self-assembly on an active metal15 and its photocatalytic reduction under laser irradiation.52–54 Importantly, photoreduced graphene oxide serves as an effective electron mediator for water splitting under visible light irradiation.16,55,56 Expanding on these insights, we propose a distinctive perspective based on previously proposed schemes and our experimental results: the spontaneous reduction of GO is a water-mediated, self-photocatalytic process under visible light. Illumination generates hole–electron pairs in GO's sp2 domains, leaving hydroxyl groups (structure 1 in Fig. 6b) as intermediates on the GO plane. The gold substrate facilitates electron transfer to the doped domain. Subsequent electron transfer reduces epoxy groups to hydroxyl, producing radical intermediates (structure 2) and OH−. These radicals, in conjunction with hydroxyl groups, further react with holes and water, forming additional hydroxyls. A GO fragment containing a 1,2-diol (structure 3) is then reduced, restoring a carbon sp2 conjugated domain, shown as structure 4 in Fig. 6b.
The hydrophilic nature of GO allows efficient noncovalent intercalation of moisture between layers.57 This suggests that the reduction process can occur more efficiently between layers, with a faster interlayer reduction in folded regions compared to flat regions. Furthermore, closely situated hydroxyl groups can undergo simultaneous reduction to form lower energy structures, producing CO2 and causing disruption in the carbon lattice, along with deoxygenation (Fig. 6 structure 5). These variations in reduction rate and mechanism between interlayer and intralayer reduction would contribute to the observed spectral features in the folded and flat regions of spontaneously reduced GO.
High-resolution chemical mapping through Tip-Enhanced Raman Scattering (TERS) offers valuable insight into the structure of the final reduced GO form and its potential configuration, and thereby the mechanisms of reduction. Nevertheless, to gain a comprehensive understanding of intermediates, such as radicals or electron–hole pairs, and to elucidate the precise pathway of GO's spontanous reduction process, the use of time-resolved spectroscopy becomes imperative.
In this study, we used correlated analysis of hyperspectral TERS imaging to investigate the spontaneous reduction of graphene oxide (GO) under ambient conditions. We obtained TERS maps on fresh and spontaneously reduced GO at two time points. D-band shifting associated with functional group variations within a rGO flake was observed for the first time. By combining TERS mapping with DFT calculations, we were able to associate the observed spectral distributions with chemical and structural disorder within reduced GO. Based on the uneven distribution of epoxy groups and carbon defects, we proposed a mechanism for the self-reduction behavior of GO, which involves a water-mediated self-photocatalytic reduction and deoxygenation process, where GO serves as both reactant and photocatalyst. The multiple reaction pathways of reduction were proposed based on the spatial distribution of epoxy groups and carbon defects, as well as the structural changes that occur during reduction. Despite proposing reduction pathways and explaining their end products, uncertainties remain about all intermediates involved in the reactions, which would be challenging to address in these studies. Identifying the intermediates and comprehensively characterizing the reduction mechanisms would necessitate higher time-resolution spectroscopy studies capable of capturing hydroxyl free radicals (HO˙), hydroxide ions OH−, and other unstable intermediates. Understanding the reduction mechanisms of GO is potentially beneficial to identifying the optimal conditions for controlled reduction, such as the type of solvent, the pH, and the temperature. This knowledge can then be used to improve the quality and uniformity of reduced GO, which is essential for its application in devices.
Our findings provide new insights into the spontaneous reduction of GO under ambient conditions. They also highlight the potential of TERS imaging for the investigation of disordered materials. Combined tip-enhanced Raman scattering mapping, statistical analysis, and DFT calculation of Raman modes improved the interpretation of highly-heterogeneous TERS data. With more detailed data analytical methods, it is possible to broaden the application of TERS to sample systems without well-separated spectral signatures, pushing TERS beyond structural characterization to spatially resolve local variations in chemical reactions. This form of time-dependent, nanoscale imaging can therefore elucidate reaction and degradation mechanisms in a wide variety of catalytic and electronic materials.
Footnotes |
† Electronic supplementary information (ESI) available: Further experimental details and information on analysis. Example TERS and micro-Raman spectra, and additional TERS intensity maps. See DOI: https://doi.org/10.1039/d3cp04711c |
‡ Present address: Department of Engineering, Westlake University, Hangzhou, China. |
This journal is © the Owner Societies 2024 |