Next Article in Journal
Recent Progress in Two-Dimensional MoTe2 Hetero-Phase Homojunctions
Previous Article in Journal
Phase Structure and Electrical Properties of Sm-Doped BiFe0.98Mn0.02O3 Thin Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Graphene Growth Directly on SiO2/Si by Hot Filament Chemical Vapor Deposition

by
Sandra Rodríguez-Villanueva
1,2,*,
Frank Mendoza
3,
Alvaro A. Instan
1,
Ram S. Katiyar
1,
Brad R. Weiner
1,4 and
Gerardo Morell
1,2
1
Department of Physics, College of Natural Science, Rio Piedras Campus, University of Puerto Rico, San Juan, PR 00925, USA
2
Molecular Sciences Research Center, University of Puerto Rico, San Juan, PR 00927, USA
3
Department of Physics, College of Arts and Sciences, Mayagüez Campus, University of Puerto Rico, Mayaguez, PR 00682, USA
4
Department of Chemistry, College of Natural Science, Rio Piedras Campus, University of Puerto Rico, San Juan, PR 00925, USA
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(1), 109; https://doi.org/10.3390/nano12010109
Submission received: 5 October 2021 / Revised: 21 October 2021 / Accepted: 26 October 2021 / Published: 30 December 2021

Abstract

:
We report the first direct synthesis of graphene on SiO2/Si by hot-filament chemical vapor deposition. Graphene deposition was conducted at low pressures (35 Torr) with a mixture of methane/hydrogen and a substrate temperature of 970 °C followed by spontaneous cooling to room temperature. A thin copper-strip was deposited in the middle of the SiO2/Si substrate as catalytic material. Raman spectroscopy mapping and atomic force microscopy measurements indicate the growth of few-layers of graphene over the entire SiO2/Si substrate, far beyond the thin copper-strip, while X-ray photoelectron spectroscopy and energy-dispersive X-ray spectroscopy showed negligible amounts of copper next to the initially deposited strip. The scale of the graphene nanocrystal was estimated by Raman spectroscopy and scanning electron microscopy.

1. Introduction

Since graphene was first obtained by microexfoliation of graphite in 2004 [1] it has been regarded as a promising material due to its excellent properties and potential applications [2,3,4,5,6,7]. Graphene’s high electron mobility, conductivity, and optical properties open up the possibility for high-speed electronics such as ultra-thin transistors, photodetectors, and optical modulators [2,3]. These attributes also contribute to the advancement of circuit boards, display panels, and solar cell technology [2,3], while its high internal surface area, electrochemical reactivities and mechanical (high stiffness and low density) properties allow greater efficiency in supercapacitors, electrochemical systems, and strain sensors, respectively [3,4,5]. Many studies have focused on obtaining graphene using a wide variety of methods [8,9,10,11], e.g., the microexfoliation of graphite [1], graphene oxide reduction [12], epitaxial growth on SiC [13,14] and chemical vapor deposition (CVD) on different substrates [15,16]. This last method is the most promising because the growth parameters can be controlled to modify the structural characteristics of the material and the number of graphene layers (monolayer, bilayer, few layers, and multilayers) deposited [17,18]. Graphene growth by CVD on metallic substrates has been used extensively, because the catalytic properties of the substrate result in a large area of high-quality graphene [19,20,21]. In order to scale this technology to industrial production, hot filament chemical vapor deposition (HFCVD) promises to be one of the leading potential techniques. This method obtains large area, high quality graphene on copper substrates with controllable growth parameters [18,22,23,24]. The hot filament dissociates the hydrogen and methane, producing active radicals that reduce the amorphous carbon to improve the quality of the graphene film [24]. The turbulent flow produced by the vertical introduction of the gas in the HFCVD provides an additional advantage for scale-up compared with the laminar flow of a tube furnace CVD. However, for use in electronic applications, current chemical vapor deposition methods require the transfer of the graphene film from the metal substrate to the dielectric, which has several drawbacks, i.e., residual chemical contamination and the risk of wrinkling or breakage of the graphene film [25]. To avoid this difficult transfer process, researchers have sought to develop new methodologies to deposit graphene directly onto non-metallic substrates such as SiO2/Si, quartz, fused silica, and others [26]. To date, there are no reports in the literature of the direct deposition of graphene on SiO2 by HFCVD, although several attempts by tube furnace CVD have been published. Table 1 presents the different methods of graphene deposition on SiO2 by other types of CVD. This table summarizes these methodologies under two classifications: Catalyst-free and metal catalyzed direct growth CVD, where both regular and plasma enhanced CVD (PECVD) are used [26].
In the first methodology (catalyst-free), the majority of the graphene growth experiments on non-metallic substrates are conducted at high temperatures (1060–1100 °C) over a long deposition time [26]. Liu et al. obtained high-quality monolayer, bilayer and few-layer graphene without any catalyst over a temperature range of 1060–1100 °C at atmospheric pressure and using methane as the carbon source [27]. Sun et al. were able to grow continuous nanocrystalline graphene at 1000 °C with good electrical properties, such as sheet resistance and Hall mobility [28]. Medina and coworkers reported that the PECVD catalyst-free growth temperature can be reduced by directly growing a nanographene film on SiO2 at low temperatures (400 °C) by using the electron cyclotron resonance CVD (ECR-CVD) method [29].
In the metal-catalyzed direct growth method, many experiments have used a sacrificial metal layer to stimulate graphene growth. McNerny et al. deposited a nickel layer on SiO2/Si wafers as a catalyst, which was subsequently delaminated using adhesive tape, leaving behind the graphene layer on the substrate [30]. They obtained a continuous (>90% coverage) graphene film on the centimeter scale, consisting of micrometer order domains and ranging from monolayer to multilayer [30]. Dong et al. deposited a copper layer on SiO2/Si substrate to synthesize graphene using a CVD tube furnace [31]. They concluded that the copper evaporation occurred after the graphene deposition, but they observed some defects and residual copper in the graphene layer, which they removed by using an FeCl3 solution [31]. Similarly, Ismach et al. deposited a copper layer on a variety of substrates (quartz, sapphire, fused silica, and SiO2/Si) to promote graphene growth [32]. They found that the copper layer was dewed and evaporated during or after graphene deposition producing areas free of copper, but residues remained all over the substrate [32]. Kato et al. combined the metal catalytic method with rapid heating plasma CVD to obtain graphene on SiO2/Si [33]. They deposited a nickel film on the substrate and using a growth temperature ranging from 600–950 °C, obtained high-quality single-layer graphene sheets with hexagonal domains, suitable for the fabrication of a graphene-based field effect transistor [33].
This paper reports a novel method suitable for industrial scale-up production to directly grow high-quality graphene on SiO2/Si substrates by HFCVD. This technique allows the deposition of graphene over the entire substrate by using the metal-catalyzed method in a limited manner. A thin copper-strip was deposited on the middle of the SiO2/Si substrate allowing the methane dehydrogenation and the carbon absorption to occur and leaving the rest of the surface free of metal. Structural, morphological, and compositional analyses were made on the graphene grown on the SiO2/Si in areas on top of and next to the copper strip. This research targets SiO2/Si substrates due to their ubiquity in graphene applications, such as photodetectors, gas sensors, solar energy, and others [3]. In addition, we use a HFCVD system that has unique advantages in terms of scalability for deposition over large area substrates [34].

2. Materials and Methods

2.1. Substrate Preparation

Nanocrystalline graphene films were grown on p-type SiO2/Si wafers with a top oxide layer of 285 nm and a thickness of 500 ± 25 µm manufactured by Graphene Supermarket (Ronkonkoma, NY, USA; https://graphene-supermarket.com/, accessed on 26 October 2021). These wafers were cut into 2 × 2 cm pieces and cleaned with: deionized water, trichlorethylene, acetone (histology grade), and isopropanol (histology grade); the last three reagents were obtained from Fisher Scientific (Pittsburgh, PA; https://www.fishersci.com/, accessed on 26 October 2021). A mixture of sulfuric acid (H2SO4 purity range of 95–98%) and hydrogen peroxide (H2O2 solution at 30% w/w in H2O), both provided by Sigma Aldrich (St. Louis, MO, USA; https://www.sigmaaldrich.com/, accessed on 26 October 2021), was prepared for a further cleaning of the substrate. A thin copper-strip (3 mm width) was deposited in the middle of the SiO2/Si substrate by sputtering (AMNPS-1 plasma-therm, Varian, Saint Petersburg, FL, USA) with a deposition time of 1 minute (cf. Figure 1). The copper target (99.99% pure) used for the deposition was obtained from the CERAC company. The thickness of the deposited copper layer was between 100–150 nm and was measured using an Ambios Technology XP-200 profilometer (Santa Cruz, CA, USA).

2.2. Graphene Synthesis

A commercial HFCVD instrument (BWS-HFCVD1000, Blue Wave, Baltimore, MD, USA; https://www.bluewavesemi.com/ accessed on 26 October 2021) was used for the graphene deposition. The reactor consists of a heated substrate holder that is positioned below three heated filaments of rhenium. The gases enter the chamber from the top with a shower-like turbulent flow, (cf. Figure 1). The HFCVD instrument allows systematic adjustment of the growth parameters e.g., pressure, gas flow rates, deposition time, substrate-to-filament distance (5–15 mm), substrate temperature and filament temperature. The SiO2/Si substrates (4 cm2) with the thin copper-strip (0.3 cm × 2.0 cm) were submitted to the graphene synthesis procedure at different growth parameters. The substrate was placed in the HFCVD as shown in Figure 1, with the copper strip perpendicular with respect to the filament orientation. The pressure and heating rate were fixed at 35 Torr and 35 °C/min, respectively, for the complete process (annealing and growth steps). During the annealing stage, the substrate was kept at 975 °C with 80 sccm of hydrogen and 20 sccm of argon for 30 min.
For the growth stage, the substrate temperature was reduced to 900 °C, and the filaments were turned on at a temperature range of 1800 °C–2300 °C in an atmosphere of methane (1–10 sccm) and hydrogen (10–50 sccm) for 30 to 120 min. Finally, the samples were cooled by spontaneous convection to room temperature. As a control study, SiO2/Si substrates without a copper-strip were also submitted to the graphene growth procedure.

2.3. Characterization

The structural characterization of graphene was conducted by Raman spectroscopy (Thermo Scientific DXR, Waltham, MA, USA) equipped with an excitation laser operating at 532 nm. The spectra were collected over a frequency range of 1100 to 3100 cm−1 with a spot size of 0.7 μm. In addition, Raman mappings were taken over an area of 150 × 100 μm2 and a step size of 2 μm; the collecting time for each point in the Raman mappings was 20 s. A morphological study of the synthesized graphene was done using a scanning electron microscope, SEM (JSM 6480LV, JEOL, Peabody, MA, USA; https://www.jeol.co.jp/en/ accessed on 26 October 2021) at different magnifications (5000×, 25,000× and 140,000×) and an atomic force microscope, AFM (Nanoscope V, Vecco, Plainview, NY, USA; https://www.veeco.com/ accessed on 26 October 2021) in tapping mode over an area of 3 × 3 µm. Compositional analyses of the graphene samples were done by energy-dispersive X-ray spectroscopy, EDS (JEOL JSM 6480LV) and X-ray photoelectron spectroscopy, XPS (PHI 5600 Physical Electronics, Chanhassen, MN, USA; https://www.phi.com/index.html accessed on 26 October 2021) over an energy range of 0 to 1200 eV.

3. Results

A structural (Raman), morphological (SEM and AFM) and compositional (EDS and XPS) analysis was done on the synthesized graphene, both on top of and next to the copper- strip deposited in the SiO2/Si substrate.

3.1. Raman Analysis

Characteristic of the Raman effect in graphene, the G peak is sensitive to sp2 carbon atoms, the 2D peak appears in response to a two-phonon vibrational process and the D peak is activated by the edges or defects in graphene [35]. All three graphene peaks were observed in the Raman spectra (cf. Figure 2), both on top of and next to the copper-strip areas on SiO2/Si substrate. For the control samples without a copper strip, these graphene peaks were not observed, indicating that the copper metal is necessary for the growth of graphene under our experimental conditions. Figure 2a,b show the Raman spectra on top of and next to the copper-strip area deposited on SiO2/Si substrate, respectively. The red and green spectra show two different signals next to the copper strip (Figure 2a) and the blue and black represent the same, but on top of the metal strip (Figure 2b). The insets show the optical images of both areas, respectively.
The G peak at 1579 cm−1, the 2D peak at 2692 cm−1 and a high D peak at 1348 cm−1 were observed in the Raman spectra for both areas. In addition, a peak at 1620 cm−1 known as D’ was found, which is related to the defects in the graphene film structure [36,37]. The D’ peak was bigger in the graphene grown on top of the copper-strip than the next to the metal film, where the peak was almost indistinguishable. This suggests that the graphene film grown on top of the copper strip has more defects. The high intensity of the D peak in both areas indicates that the carbon films are composed of nanometer-scale crystallites [36]. The presence of this peak (D) could also be associated with defects in the crystallite structure [18,37,38].
The average intensity ratio between the D and G peaks (ID/G) yields an estimate of the graphene grain size [39,40] and the level of the defective crystallites [36,37,41,42]. In our case, these values were between 0.30 ± 0.04 and 0.80 ± 0.03 next to the copper strip. The higher ID/G values, 0.45 ± 0.07 and 0.87 ± 0.03, were found on top of the metal strip. Although, we had a significant observed D peak, the average of the full width at half maximum (FWHM) of the D, G and 2D peaks indicates good quality crystallites [36]. The FWHM of these peaks on top of the copper strip were 35 ± 1 cm−1, 25 ± 1 cm−1, and 56 ± 3 cm−1, respectively and in areas next to the copper strip were: 38 ± 2 cm−1, 29 ± 1 cm−1 and 52 ± 2 cm−1.
To calculate the crystal size from the Raman data, we employ the Cancado equation (Equation (1)) [38], where La corresponds to the crystallite size, λl represents the wavelength of the excitation laser, ID/IG is the intensity ratio of the D and G peaks and 2.4 × 10 −10 is the proportionality constant between ID/IG and Lα. We found that the Lα on top of and next to the copper strip was in the range of 24.03 to 64.07 nm and 22.11 to 42.72 nm, respectively, in agreement with the D peak characteristics associated to the nanocrystals, but different from the grain size (35–140 nm) measured by SEM (vide infra):
L α ( nm ) = ( 2.4 × 10 10 )   λ l 4   ( I D / I G ) 1
The difference in the particle size estimates is likely due to the multiple phonon dispersion produced by defects inside of the graphene crystallites [37,43]. These imperfections in the crystal affect the intensity ratio between the D and G peaks in the Raman spectra, resulting in false behavior of smaller grains [37,43]. To estimate the contribution of these defects, we use Equation (2) [44,45], where LD represents the inter-defect distance, EL is the excitation energy and the defect concentration corresponds to 1/ L D 2 [45]. Our results of the average LD in areas next to and on top of copper strip were 18 nm and 10 nm, respectively. We also estimate the defect concentration for both areas, next to and on top of copper strip with values of 3 × 10−3/nm 2 and 7 × 10−3/nm 2, respectively. These results confirm that some point defects are present in the nanocrystals and contribute to the ID/IG ratio intensity. In addition, we corroborate that higher concentration of defective crystals are present on top of the copper strip versus next to this metal film:
L D 2   ( nm 2 ) = 3600 E L 4 ( I D / I G ) 1
The G and 2D peaks characteristically correspond to the signal for graphitic materials [18], where the intensity of these peaks was higher on top of the copper-strip areas than next to this film.
Raman mapping (cf. Figure 3) was done to understand the uniformity of graphene layers on the SiO2/Si substrate and to estimate the number of graphene layers through the intensity ratio of the 2D/G peaks [18,39]. In Figure 3a,b, a visual image of the graphene growth is shown next to and on top of the copper-strip for a selected mapping area of 150 × 100 μm2. In Figure 3a, it is possible to identify the general uniformity of the graphene growth throughout the mapped areas, while in Figure 3b the presence of the copper particles are clearly observed. Figure 3c,d show the Raman mapping of the intensity ratio of 2D/G peaks, for the same areas next to and on top of the copper-strip shown in Figure 3a,b. The average 2D/G ratio was 0.70 ± 0.05 and 0.50 ± 0.07 for Figure 3c,d, respectively. It is possible to estimate the number of graphene layers from the value of the 2D/G intensity ratio, which in our case corresponds to few layers of graphene [18,32,35]. However, other reasons such as the doping levels in the graphene layer can have an effect on this value (2D/G intensity), leading to an incorrect estimate of the number of layers [44].

3.2. SEM Analysis

Figure 4a,b show the SEM images taken in two areas next to the copper strip with a magnification of 140,000×. Figure 4a shows an area 8 mm from the copper film, while Figure 4b is an area closer (4 mm) to the copper strip. Similarly, Figure 4c,d show two different areas on top of the copper strip, upper and middle.
From the SEM images, it was possible to estimate the size of the graphene crystals from the scale bar to ca.100 nm. By measuring many crystals, we obtained an average size of 120 nm and a range of 100 to 140 nm for particles next to the copper strip, and smaller particles (35–120 nm; average size = 74 nm) on top of the copper-strip. At lower magnification (5000×), no copper particles were observed next to the copper film.

3.3. AFM Analysis

Figure 5 shows the AFM measurements for graphene growth on SiO2/Si substrate for both next to (cf. Figure 5a) and on top of (cf. Figure 5b) the copper-strip area, respectively. The copper grains were identified with an average height of 50 nm (Figure 5b) and uniform graphene layers were observed next to the copper strip with an average height of 5 nm (Figure 5a) corresponding to 6–12 graphene layers [18,46,47,48], supporting our calculations obtained from the Raman spectra. A nanocrystalline pattern was expected to be found, [36] however this was not identified because the deposited carbon material was composed of more than one layer of graphene. Nevertheless, two different morphologies were observed between areas on top of and next to the copper-strip.

3.4. EDS Analysis

A compositional analysis of graphene on SiO2/Si samples was done by EDS. In areas next to the copper strip (cf. Figure 6a), the following elements were identified (with their respective atomic concentrations): silicon (77.28%), oxygen (19.37%) and carbon (3.34%). In the EDS spectra on top of the copper-strip, the following elements were observed, silicon (57.02%), oxygen (11.89%), copper (20.08%) and carbon (11.01%) (cf. Figure 6b).
These atomic concentrations are consistent with the 2D/G intensity ratio in the Raman mapping experiment, where the lower values were found on top of the copper strip areas, indicating that more carbon atoms were deposited [35]. Although a higher carbon concentration was presented on top of the copper strip, a considerable percentage next to the metal film was identified. Additionally, no trace of copper was found next to the copper strip area, showing that there is graphene growth in metal-free areas.

3.5. XPS Analysis

XPS measurements were taken both next to and on top of the copper-strip. Figure 7a,d show the spectra of the elements found in both areas, respectively. The carbon 1s (C1s) peak was observed in both areas (Figure 7b,e). The raw data is shown on the dotted line and the solid lines represent the contribution of all the peaks after deconvolution. Contribution peaks were observed at 284.6 eV, 285.9 eV and 290.0 eV, corresponding to C-C, C-O and O-C=O respectively [38,49,50]. The presence of oxygen is confirmed in both areas in the XPS spectra (Figure 7a,d). The incorporation of oxygen most likely occurred after the graphene growth following exposure to air. The copper peaks (Cu 2p3/2: 930–937 eV and Cu 2p1/2: −954 eV) were observed on top of the copper-strip (Figure 7f), as expected. However, this metal shows a very small signal next to the copper-strip area (Figure 7c). Signals from other metals such as Fe (Fe 2p3/2: 706.7–710.9 eV), Co (Co 2p3/2: 778.1–780.2 eV) and Ni (Ni 2p3/2: 852.5–854.4 eV) were not observed on areas next to and on top of the copper-strip. The absence of other metals demonstrates that the graphene growth was either catalyst free or catalyzed by copper [25]. (Figure 7a,d).
The structural (Raman), morphological (SEM and AFM) and compositional (EDS and XPS) characteristics of the graphene on SiO2/Si substrate samples were measured. This characterization confirmed that this graphitic material grew over all areas of the SiO2 substrate at the nanocrystalline scale. The calculated grain size from Raman measurements was between 24.03 to 64.07 nm (next to the copper-strip); however, defects in the crystal due to phonon scattering may lead to an error in this estimate. These defects inside of the graphene nanocrystal were corroborated by the calculation of the inter-distance defect (Equation (2)).
The real size was confirmed through the images taken by the SEM technique where the particle size was in a range of 35 to 140 nm with an average of 120 nm (next to the copper-strip). The growth mechanism most likely begins with dehydrogenation of methane by the hot filament [18]. In the absence of copper, no graphene is observed, and therefore the growth must be catalyzed by the metal. This raises the question of whether the graphene is catalyzed on the metal film and migrates across the surface to cover the substrate (Figure 8a), or if the catalysis occurs due to vapor phase copper species above the surface (Figure 8b) [18,32,51,52,53]. If the vapor phase metal-catalyzed mechanism is operative, the expectation is that copper should be present across the substrate. While we do not see abundant amounts of copper next to the copper film, we cannot conclusively rule out the mechanism shown in Figure 8b because of the small signal observed in our XPS data. According to the growth distribution of graphene on the substrate we suggest that some crystals grew as migration from the copper-strip (Figure 8a), but some of the crystals next to the metal film were formed by the copper vapor catalyst effect (Figure 8b) [51,52] that is evaporated during the growth stage [31,51,52,53], leaving a small residual amount consistent with our XPS data.

4. Discussion

This study demonstrates, for the first time, a method to deposit polycrystalline graphene directly onto SiO2/Si by HFCVD, avoiding a complex graphene transfer process. In this method, a thin copper-strip of 0.3 cm × 2.0 cm was deposited in the middle of a 4 cm2 substrate, leaving most of the substrate surface free of this metal. The structural analysis was done by Raman spectra to verify the graphene growth characteristics. SEM and AFM images allowed us to determine the graphene’s topography on the SiO2/Si substrate. Additionally, copper residues were observed on top of the copper-strip areas, but these were not present in areas next to the metal. A compositional study was made through EDS and XPS measurements, indicating the presence of carbon in all samples and the virtual absence of copper in areas next to the metal-strip. This work demonstrates that the thin copper-strip deposited on the middle of the SiO2/Si enables the graphene growth over all the substrate. By eliminating the need for a mechanical transfer step in the device fabrication process, this accomplishment opens up the possibility of integrating graphene with currently available silicon device technologies. Further research, needed to continuously improve the quality of the graphene deposition, is ongoing in our laboratories. One approach is the reduction of the nucleation density [15,54,55,56] by modifying the methane and hydrogen gas flow rates that will allow an increment in the graphene crystal size and reduction of the point defects [55,56].

5. Conclusions

This work presents an approach to directly grow graphene on SiO2/Si by HFCVD, using the metal catalyzed method in a limited manner. The crystal size, structure, and inter-defect distance of the nanocrystalline graphene were estimated by SEM, AFM, and Raman measurements, respectively. EDS and XPS analyses confirmed the presence of graphene on SiO2/Si with negligible amount of copper in the area next to the copper strip. Our study allows the possibility of growing graphene directly on dielectrics without a transfer process and the opportunity to produce it on an industrial scale.

Author Contributions

Conceptualization, S.R.-V., F.M., B.R.W. and G.M.; Methodology, S.R.-V., F.M., B.R.W. and G.M.; Validation, S.R.-V., R.S.K. and A.A.I.; Formal Analysis, S.R.-V. and F.M.; Investigation, S.R.-V. and A.A.I.; Resources, R.S.K., B.R.W. and G.M.; Data Curation, S.R.-V.; Writing—Original Draft Preparation, S.R.-V.; Writing-Review & Editing, B.R.W. and G.M.; Visualization, S.R.-V. and F.M.; Supervision, B.R.W. and G.M.; Project Administration, B.R.W. and G.M.; Funding Acquisition, B.R.W. and G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded in part by NSF EPSCoR CAWT, grant number OIA-1849243, PR Space Grant, grant number 80NSSC20M0052 and PR NASA EPSCoR grant number 80NSSC19M0049.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data can be obtained from the corresponding author.

Acknowledgments

The authors acknowledge the Materials Characterization Center (MCC) for providing access to Raman and SEM facilities and to Luis Fonseca for access to the sputtering equipment.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Novoselov, K.; Geim, A.; Morozov, S.; Jiang, D.; Zhang, Y.; Dubonos, S.; Grigorieva, I.; Firsov, A. Electric field in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. De Fazio, D.; Purdie, D.; Ott, A.; Braeuninger-Weimer, P.; Khodkov, T.; Goossens, S.; Taniguchi, T.; Watanabe, K.; Livreri, P.; Koppens, F.; et al. High-mobility, wet-transferred graphene grown by chemical vapor deposition. ACS Nano 2019, 13, 8926–8935. [Google Scholar] [CrossRef] [Green Version]
  3. Randviir, E.; Brownson, D.; Banks, C. A decade of graphene research: Production, applications and outlook. Mater. Today 2014, 17, 426–432. [Google Scholar] [CrossRef]
  4. Bunch, J.; van der Zande, A.; Verbridge, S.; Frank, I.; Tanenbaum, D.; Parpia, J.; Craighead, H.; McEuen, P. Electromechanical resonators from graphene sheets. Science 2007, 315, 490–493. [Google Scholar] [CrossRef] [Green Version]
  5. Raju, A.; Lewis, A.; Derby, B.; Young, R.; Kinloch, I.; Zan, R.; Novoselov, K. Wide-area strain sensors based upon graphene-polymer composite coatings probed by Raman spectroscopy. Adv. Func. Mater. 2014, 24, 2865–2874. [Google Scholar] [CrossRef]
  6. Kong, W.; Kum, H.; Bae, S.; Shim, J.; Kim, H.; Kong, L.; Meng, Y.; Wang, K.; Kim, C.; Kim, J. Path towards graphene commercialization from lab to market. Nat. Nanotechnol. 2019, 14, 927–938. [Google Scholar] [CrossRef]
  7. Edwards, R.; Coleman, K. Graphene synthesis: Relationship to applications. Nanoscale 2013, 5, 38–51. [Google Scholar] [CrossRef]
  8. Bonaccorso, F.; Lombardo, A.; Hasan, T.; Sun, Z.; Colombo, L.; Ferrari, A. Production and processing of graphene and 2D crystals. Mater. Today 2012, 15, 564–589. [Google Scholar] [CrossRef]
  9. Tan, H.; Wang, D.; Guo, Y. Thermal growth of graphene: A review. Coatings 2018, 8, 40. [Google Scholar] [CrossRef] [Green Version]
  10. Cooper, D.; D’Anjou, B.; Ghattamaneni, N.; Harack, B.; Hilke, M.; Horth, A.; Majlis, N.; Massicotte, M.; Vandsburger, L.; Whiteway, E.; et al. Experimental review of graphene. Condens. Matter Phys. 2012, 2012, 501686. [Google Scholar] [CrossRef] [Green Version]
  11. Woehrl, N.; Ochedowski, O.; Gottlieb, S.; Shibasaki, K.; Schulz, S. Plasma-enhanced chemical vapor deposition of graphene on copper substrates. AIP Adv. 2014, 4, 047128. [Google Scholar] [CrossRef] [Green Version]
  12. Pei, S.; Cheng, H. The reduction of graphene oxide. Carbon 2012, 50, 3210–3228. [Google Scholar] [CrossRef]
  13. Mishra, N.; Boeckl, J.; Motta, N.; Iacopi, F. Graphene growth on silicon carbide: A review. Phys. Status Solidi A 2016, 213, 2277–2289. [Google Scholar] [CrossRef]
  14. Reza, G.; Iakimov, T.; Yakimova, R. Epitaxial Graphene on SiC: A Review of Growth and Characterization. Crystals 2016, 6, 53. [Google Scholar]
  15. Antonova, I. Chemical vapor deposition growth of graphene on copper substrates: Current trends. Phys. Uspekhi 2013, 56, 1013–1020. [Google Scholar] [CrossRef]
  16. Mattevi, C.; Kima, H.; Chhowalla, M. A review of chemical vapour deposition of graphene on copper. J. Mater. Chem. 2011, 21, 3324–3334. [Google Scholar] [CrossRef]
  17. Umair, A.; Raza, H. Controlled synthesis of bilayer Graphene on nickel. Nanoscale Res. Lett. 2012, 7, 437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Limbu, T.; Hernández, J.; Mendoza, F.; Katiyar, R.; Razink, J.; Makarov, V.; Weiner, B.; Morell, G. A novel approach to the layer-number-controlled and grain-size-controlled growth of high-quality graphene for nanoelectronics. Appl. Nano Mater. 2018, 1, 1502–1512. [Google Scholar] [CrossRef]
  19. Li, X.; Cai, W.; An, J.; Kim, S.; Nah, J.; Yang, D.; Piner, R.; Velamakanni, A.; Jung, I.; Tutuc, E.; et al. Large-area synthesis of high-quality and uniform graphene films on copper foils. Science 2009, 324, 1312–1314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Li, X.; Magnuson, C.; Venugopal, A.; An, J.; Suk, J.; Han, B.; Borysiak, M.; Cai, W.; Velamakanni, A.; Zhu, Y.; et al. Graphene films with large domain size by a two-step chemical vapor deposition process. Nano Lett. 2010, 10, 4328–4334. [Google Scholar] [CrossRef] [Green Version]
  21. Petrone, N.; Dean, C.; Meric, I.; van der Zande, A.; Huang, P.; Wang, L.; Muller, D.; Shepard, K.; Hone, J. Chemical vapor deposition-derived graphene with electrical performance of exfoliated graphene. Nano Lett. 2012, 12, 2751–2756. [Google Scholar] [CrossRef]
  22. Mendoza, F.; Limbu, T.; Weiner, B.; Morell, G. Large-area bilayer graphene synthesis in the hot filament chemical vapor deposition reactor. Diam. Relat. Mater. 2015, 51, 34–38. [Google Scholar] [CrossRef]
  23. Lau, K.; Caulfield, J.; Gleason, K. Structure and morphology of fluorocarbon films grown by hot filament chemical vapor deposition. Chem. Mater. 2000, 12, 3032–3037. [Google Scholar] [CrossRef]
  24. Hafiz, S.; Chong, S.; Huang, N.; Rahman, S. Fabrication of high-quality graphene by hot-filament thermal chemical vapor deposition. Carbon 2015, 86, 1–11. [Google Scholar] [CrossRef]
  25. Chen, J.; Wen, Y.; Guo, Y.; Wu, B.; Huang, L.; Xue, Y.; Geng, D.; Wang, D.; Yu, G.; Liu, Y. Oxygen-Aided synthesis of polycrystalline graphene on silicon dioxide substrates. J. Am. Chem. Soc. 2011, 133, 17548–17551. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, H.; Yu, G. Direct CVD graphene growth on semiconductors and dielectrics for transfer-free device fabrication. Adv. Mater. 2016, 28, 4956–4975. [Google Scholar] [CrossRef]
  27. Liu, Q.; Gong, Y.; Wang, T.; Chan, W.; Wu, J. Metal-catalyst-free and controllable growth of high-quality monolayer and AB-stacked bilayer graphene on silicon dioxide. Carbon 2016, 96, 203–211. [Google Scholar] [CrossRef] [Green Version]
  28. Sun, J.; Lindvall, N.; Cole, M.; Wang, T.; Booth, T.; Bøggild, P.; Teo, K.; Liu, J.; Yurgens, A. Controllable chemical vapor deposition of large area uniform nanocrystalline graphene directly on silicon dioxide. Int. J. Appl. Phys. 2012, 111, 044103. [Google Scholar]
  29. Medina, H.; Lin, Y.; Jin, C.; Lu, C.; Yeh, C.; Huang, K.; Suenaga, K.; Robertson, J.; Chiu, P. Metal-Free Growth of Nanographene on Silicon Oxides for Transparent Conducting Applications. Adv. Funct. Mater. 2012, 22, 2123–2128. [Google Scholar] [CrossRef]
  30. McNerny, D.; Viswanath, B.; Copic, D.; Laye, F.; Prohoda, C.; Brieland, A.; Polsen, E.; Dee, N.; Veerasamy, V.; Hart, A. Direct fabrication of graphene on SiO2 enabled by thin film stress engineering. Sci. Rep. 2014, 4, 5049. [Google Scholar] [CrossRef] [Green Version]
  31. Dong, Y.; Xie, Y.; Xu, C.; Li, X.; Deng, J.; Fan, X.; Pan, G.; Wang, Q.; Xiong, F.; Fu, Y.; et al. Transfer-free, lithography-free, and micrometer-precision patterning of CVD graphene on SiO2 toward all-carbon electronics. APL Mater. 2018, 6, 026802. [Google Scholar] [CrossRef] [Green Version]
  32. Ismach, A.; Druzgalski, C.; Penwell, S.; Schwartzberg, A.; Zheng, M.; Javey, A.; Bokor, J.; Zhang, Y. Direct Chemical Vapor Deposition of Graphene on Dielectric Surfaces. Nano Lett. 2010, 10, 1542–1548. [Google Scholar] [CrossRef]
  33. Kato, T.; Hatakeyama, R. Direct growth of doping-density controlled hexagonal graphene on SiO2 substrate by rapid-heating plasma CVD. ACS Nano 2012, 6, 8508–8515. [Google Scholar] [CrossRef] [PubMed]
  34. Zimmer, J.; Ravi, K. Aspects of scaling CVD diamond reactors. Diam. Relat. Mater. 2006, 15, 229–233. [Google Scholar] [CrossRef]
  35. Ni, Z.; Wang, Y.; Yu, T.; Shen, Z. Raman spectroscopy and imaging of graphene. Nano Res. 2008, 1, 273–291. [Google Scholar] [CrossRef] [Green Version]
  36. Yang, W.; He, C.; Zhang, L.; Wang, Y.; Shi, Z.; Cheng, M.; Xie, G.; Wang, D.; Yang, R.; Shi, D.; et al. Growth, Characterization, and Properties of Nanographene. Small 2012, 8, 1429–1435. [Google Scholar] [CrossRef]
  37. Wu, J.; Lin, M.; Cong, X.; Liu, H.; Tan, P. Raman spectroscopy of graphene-based materials and its applications in related devices. Chem. Soc. Rev. 2018, 47, 1822–1873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Hawaldar, R.; Merino, P.; Correia, M.; Bdikin, I.; Grácio, J.; Méndez, J.; Martin, J.; Kumar, M. Large-area high-throughput synthesis of monolayer graphene sheet by Hot Filament Thermal Chemical Vapor Deposition. Sci. Rep. 2012, 2, 682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Cançado, L.; Takai, K.; Enoki, T. General equation for the determination of the crystallite size La of nanographite by Raman spectroscopy. Appl. Phys. Lett. 2006, 88, 163106. [Google Scholar] [CrossRef]
  40. Tuinstra, F.; Koenig, J. Raman Spectrum of Graphite. J. Chem. Phys. 1970, 53, 1126. [Google Scholar] [CrossRef] [Green Version]
  41. Jorio, A.; Ferreira, E.; Moutinho, M.; Stavale, F.; Achete, C.; Capaz, R. Measuring disorder in graphene with the G and D bands. Phys. Status Solidi B 2010, 247, 2980–2982. [Google Scholar] [CrossRef]
  42. Eckmann, A.; Felten, A.; Mishchenko, A.; Britnell, L.; Krupke, R.; Novoselov, K.; Casiraghi, C. Probing the Nature of Defects in Graphene by Raman Spectroscopy. Nano Lett. 2012, 12, 3925–3930. [Google Scholar] [CrossRef] [Green Version]
  43. Ferrari, A.; Basko, D. Raman spectroscopy as a versatile tool for studying the properties of graphene. Nat. Nanotechnol. 2013, 8, 235–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Rassapa, S.; Caridad, J.; Schulte, L.; Cagliani, A.; Borah, D.; Morris, M.; Bøggild, P.; Ndoni, S. High quality sub-10 nm graphene nanoribbons by on-chip PS-b-PDMS block copolymer lithography. RSC Adv. 2015, 5, 66711–66717. [Google Scholar] [CrossRef] [Green Version]
  45. Mikhailov, S. Measuring disorder in graphene with Raman spectroscopy. In Physics and Applications of Graphene—Experiments; InTech Publishers: London, UK, 2011; pp. 439–454. [Google Scholar]
  46. Reina, A.; Jia, X.; Ho, J.; Nezich, D.; Son, H.; Bulovic, V.; Dresselhaus, M.; Kong, J. Large Area, Few-Layer Graphene Films on Arbitrary Substrates by Chemical Vapor Deposition. Nano Lett. 2009, 9, 30–35. [Google Scholar] [CrossRef] [PubMed]
  47. Yao, Y.; Ren, L.; Gao, S.; Li, S. Histogram method for reliable thickness measurements of graphene films using atomic force microscopy (AFM). J. Mater. Sci. Technol. 2017, 33, 815–820. [Google Scholar] [CrossRef]
  48. Nemes-Incze, P.; Osvatha, Z.; Kamaras, K.; Biro, L. Anomalies in thickness measurements of graphene and few layer graphite crystals by tapping mode atomic force microscopy. Carbon 2008, 46, 1435–1442. [Google Scholar] [CrossRef] [Green Version]
  49. Ferrah, D.; Penuelas, J.; Bottela, C.; Grenet, G.; Ouerghi, A. X-ray photoelectron spectroscopy (XPS) and diffraction (XPD) study of a few layers of graphene on 6H-SiC (0001). Surf. Sci. 2013, 615, 47–56. [Google Scholar] [CrossRef]
  50. Moulder, J. Handbook of X-Ray Photoelectron Spectroscopy; Physical Electronics Division, Perkin-Elmer Corporation: Eden Prairie, MN, USA, 1992. [Google Scholar]
  51. Teng, P.; Lu, C.; Akiyama, K.; Lin, Y.; Yeh, C.; Suenaga, K.; Chiu, P. Remote Catalyzation for Direct Formation of Graphene Layers on Oxides. Nano Lett. 2012, 12, 1379–1384. [Google Scholar] [CrossRef]
  52. Song, Y.; Liu, J.; Quan, L.; Pan, N.; Zhu, H.; Wang, X. Size Dependence of Compressive Strain in Graphene Flakes Directly Grown on SiO2/Si Substrate. J. Phys. Chem. C 2014, 118, 12526–12531. [Google Scholar] [CrossRef]
  53. Kim, H.; Song, I.; Park, C.; Son, M.; Hong, M.; Kim, Y.; Kim, J.; Shin, H.; Baik, J.; Choi, H. Copper Vapor-Assisted Direct Growth of High Quality and Metal-Free Single Layer Graphene on Amorphous SiO2 Substrate. ACS Nano 2013, 7, 6575–6582. [Google Scholar] [CrossRef] [PubMed]
  54. Luo, B.; Caridad, J.; Whelan, P.; Thomsen, J.; Mackenzie, D.; Cabo, A.; Mahatha, S.; Bianchi, M.; Hofmann, P.; Jepsen, P.; et al. Sputtering an exterior metal coating on copper enclosure for large-scale growth of single-crystalline graphene. 2D Mater. 2017, 4, 045017. [Google Scholar] [CrossRef]
  55. Muñoz, R.; Gómez, C. Review of CVD synthesis of graphene. Chem. Vap. Depos. 2013, 19, 297–322. [Google Scholar] [CrossRef] [Green Version]
  56. Bhaviripudi, S.; Jia, X.; Dresselhaus, M.; Kong, J. Role of kinetic factors in chemical vapor deposition synthesis of uniform large area graphene using copper catalyst. Nano Lett. 2010, 10, 4128–4133. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Hot filament chemical vapor deposition (HFCVD) reactor schematic and the SiO2/Si substrate with the deposited copper-strip.
Figure 1. Hot filament chemical vapor deposition (HFCVD) reactor schematic and the SiO2/Si substrate with the deposited copper-strip.
Nanomaterials 12 00109 g001
Figure 2. Raman spectra of graphene on SiO2/Si substrates and its respective peaks (D, G and 2D): (a) Next to the copper-strip areas and (b) On top of the copper-strip areas.
Figure 2. Raman spectra of graphene on SiO2/Si substrates and its respective peaks (D, G and 2D): (a) Next to the copper-strip areas and (b) On top of the copper-strip areas.
Nanomaterials 12 00109 g002
Figure 3. Raman mapping for graphene growth on SiO2/Si substrate. Where (a,b) represent the optical images of the selected mapping area (150 × 100 μm2) next to and on top of the copper-strip, respectively. (c,d) show the ratio between the intensity of 2D/G peaks in the same areas as in (a,b).
Figure 3. Raman mapping for graphene growth on SiO2/Si substrate. Where (a,b) represent the optical images of the selected mapping area (150 × 100 μm2) next to and on top of the copper-strip, respectively. (c,d) show the ratio between the intensity of 2D/G peaks in the same areas as in (a,b).
Nanomaterials 12 00109 g003
Figure 4. SEM measurements of the graphene growth on SiO2/Si substrate: (a,b) show the SEM image taken in two areas next to the copper-strip at 140,000×. Similarly, (c,d) show two areas on top of the copper strip at the same magnification. In all cases, crossed arrows represent the position relative to the copper strip where the image was taken.
Figure 4. SEM measurements of the graphene growth on SiO2/Si substrate: (a,b) show the SEM image taken in two areas next to the copper-strip at 140,000×. Similarly, (c,d) show two areas on top of the copper strip at the same magnification. In all cases, crossed arrows represent the position relative to the copper strip where the image was taken.
Nanomaterials 12 00109 g004
Figure 5. AFM measurements of the graphene growth on SiO2/Si substrate: (a,b) show the AFM images taken next to and on top of the copper-strip, respectively.
Figure 5. AFM measurements of the graphene growth on SiO2/Si substrate: (a,b) show the AFM images taken next to and on top of the copper-strip, respectively.
Nanomaterials 12 00109 g005
Figure 6. EDS spectrum of graphene on SiO2/Si substrate (a,b) shows the EDS spectrum next to and on top of the copper-strip, respectively.
Figure 6. EDS spectrum of graphene on SiO2/Si substrate (a,b) shows the EDS spectrum next to and on top of the copper-strip, respectively.
Nanomaterials 12 00109 g006
Figure 7. XPS measurements of the graphene growth on SiO2/Si substrate: (a) shows the XPS full composition spectra, (b) carbon peak after deconvolution and (c) the copper peaks taken next to the copper-strip. Then (df) represent the same but on top of the copper-strip.
Figure 7. XPS measurements of the graphene growth on SiO2/Si substrate: (a) shows the XPS full composition spectra, (b) carbon peak after deconvolution and (c) the copper peaks taken next to the copper-strip. Then (df) represent the same but on top of the copper-strip.
Nanomaterials 12 00109 g007
Figure 8. Schematic of the graphene growth mechanism on SiO2/Si substrate. (a) graphene migration from the copper-strip film and (b) the catalytic effect of the copper vapor to form graphene. In both figures, CH4/H2 molecules pass through the hot filaments prior to deposition. For more detail on the mechanism, see text.
Figure 8. Schematic of the graphene growth mechanism on SiO2/Si substrate. (a) graphene migration from the copper-strip film and (b) the catalytic effect of the copper vapor to form graphene. In both figures, CH4/H2 molecules pass through the hot filaments prior to deposition. For more detail on the mechanism, see text.
Nanomaterials 12 00109 g008
Table 1. Methodologies to grow graphene on non-metallic substrates by CVD. Some growth parameters such as gas flow, temperature and carbon source are presented.
Table 1. Methodologies to grow graphene on non-metallic substrates by CVD. Some growth parameters such as gas flow, temperature and carbon source are presented.
MethodCVD TypeSubstratePre-Growth StepCarbon Source/
Temperature
References
Catalyst-freeTube FurnaceSiO2 (0, 90, 300,
500 nm)/Si
H2 (70–160 sccm)/1060–1100 °CCH4 (30 sccm)/1060–1100 °C[27]
SiO2 (300 nm)/SiH2 (50 sccm) and Ar (1000 sccm)/1000 °CCH4 (300 sccm)/1000 °C[28]
ECR plasmaSiO2/Si, quartz, and glassAr (5sccm)/400 °CC2H4 (0.12 sccm) and Ar (0.12 sccm)/400 °C[29]
Metal-catalyzedTube FurnaceNi layer/siliconH2 or He (400sccm)/
900 °C
CH4 or C2H2 (50 sccm) and H2 (50 sccm)/900 °C[30]
Cu layer (60 nm)/SiO2 (300 nm)/SiH2 (35 sccm)/1000 °CCH4 (30 sccm) and H2 (20 sccm)/960 °C[31]
Cu layer (450 to 100 nm)/quartz, sapphire, SiO2 (300 nm)/Si, and fused silicaH2 (35 sccm)/1000 °CCH4 (35 sccm) and H2 (2 sccm)/1000 °C [32]
Rapid heating plasmaNi film (55 nm)/SiO2 (300 nm)/SiCH4:H2 (9:1)/600–975 °C CH4:H2 (9:1)/950 °C [33]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rodríguez-Villanueva, S.; Mendoza, F.; Instan, A.A.; Katiyar, R.S.; Weiner, B.R.; Morell, G. Graphene Growth Directly on SiO2/Si by Hot Filament Chemical Vapor Deposition. Nanomaterials 2022, 12, 109. https://doi.org/10.3390/nano12010109

AMA Style

Rodríguez-Villanueva S, Mendoza F, Instan AA, Katiyar RS, Weiner BR, Morell G. Graphene Growth Directly on SiO2/Si by Hot Filament Chemical Vapor Deposition. Nanomaterials. 2022; 12(1):109. https://doi.org/10.3390/nano12010109

Chicago/Turabian Style

Rodríguez-Villanueva, Sandra, Frank Mendoza, Alvaro A. Instan, Ram S. Katiyar, Brad R. Weiner, and Gerardo Morell. 2022. "Graphene Growth Directly on SiO2/Si by Hot Filament Chemical Vapor Deposition" Nanomaterials 12, no. 1: 109. https://doi.org/10.3390/nano12010109

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