359
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Continuous flow laser-induced unzipping of multiwalled carbon nanotubes

ORCID Icon
Article: 2310885 | Received 23 Jan 2023, Accepted 23 Jan 2024, Published online: 09 Feb 2024

Abstract

Unzipped multiwalled carbon nanotubes (MWCNTs) provide a route to graphene carbon ribbons (GNRs), which have application in electronic devices. Pulsed irradiation using an Nd:YAG laser operating at 1064 nm mediates such unzipping of MWCNTs dispersed in ethanol under shear stress within the vortex fluidic device (VFD). The method is scalable with the thin film device operating under continuous flow, while also avoiding the use of harsh chemicals and auxiliary substances. Unzipped MWCNTs are formed in 90% yield and have been characterized using scanning electron microscopy, transmission electron microscopy, Raman spectroscopy, thermogravimetric analysis, X-ray powder diffraction and X-ray photoelectron spectroscopy. The systematically derived optimal laser power for unzipping was 250 mJ and increasing the laser power results in the fragmentation of the MWCNTs.

1. Introduction

Manipulation of the structure of carbon nanotubes (CNTs) in a controlled way is crucial for numerous nanotechnology applications [Citation1–5]. In this context, unzipping of multiwalled CNTs (MWCNTs) has been extensively studied in recent years due to their attractive electrical and mechanical properties [Citation1,Citation2,Citation6,Citation7]. Indeed, unzipping tubular MWCNTs is an effective strategy for synthesizing graphene nanoribbons (GNRs)/graphene material in large quantities [Citation8,Citation9] with motivation for this work primarily a desire to create flat carbon networks. The initial diameter and length of CNTs can determine the size of the fabricated GNRs, noting that flat semiconducting GNRs with a width of <10 nm are desirable, in particular, in the production of electronic devices [Citation10–12].

Several “top-down and bottom-up” methods for producing GNRs have been reported, including chemical vapour deposition, using plasmas, lithography, chemical solution-based oxidative processes and laser ablation [Citation13–22]. The disadvantages of these technologies include low yield of production, complex processing step, time-consuming operations and the use of toxic auxiliary chemicals [Citation20,Citation23–27]. Rao et al. used a fast and simple method based on laser irradiation to unzip MWCNTs on a glass substrate [Citation28]. However, this method is not scalable, has limited productivity, and it is inherently difficult to collect the unzipped material. Kumar et al. demonstrated that laser irradiation can unzip MWCNTs into GNRs dispersed in dimethylformamide (DMF) [Citation24], and Shim et al. have developed a simple method for fabricating GNRs involving laser irradiation of a solution of MWCNTs in ethanol [Citation29]. However, this method was preformed using batch processing and the product was not uniform.

GNRs and their composite can be used as shielding materials for electromagnetic interference (EMI) [Citation30,Citation31]. Park et al. [Citation32] synthesized graphene nanoribbons (GNRs) by unzipping MWCNTs and incorporated the GNRs into thermoplastic polyurethane (TPU) and fabricated a composite material for EMI shielding. These composite materials (GNRs/TPU) achieved 24.9 dB EMI shielding effectiveness [Citation32]. Sun et al. [Citation33] fabricated graphene on polyimide (PI) substrates using laser scribing. The fabricated materials showed an EMI shielding effectiveness between ∼20 dB and ∼50 dB, depending on the nature of the graphene layers [Citation33].

Recently, we reported the unzipping of MWCNTs, in moderate yield (75%) using a vortex fluidic device (VFD) operating under continuous flow [Citation34]. Here the MWCNTs are suspended in aqueous hydrogen peroxide (H2O2) and processed under high shear continuous flow conditions, resulting in partially and fully unzipped MWCNTs. [Citation34]. We note that laser irradiation of an aqueous hydrogen peroxide suspension of MWCNTs in a VFD results in the formation of carbon dots [Citation35]. The VFD itself is an inexpensive thin film microfluidic platform housing a tilted, rapidly rotating quartz tube, typically 18.5 cm in length and 20 mm in outside diameter. VFD processing induces mechanical energy, as well as intense micromixing within the liquid in a controlled way [Citation36–38], with processing using the VFD suited for developing processes high in green chemistry metrics [Citation36,Citation37]. The VFD has a diverse range of applications in general, also including slicing carbon nanotubes (CNTs) [Citation39,Citation40] exfoliating 2D materials [Citation41–43], controlling carbon nanomaterial morphologies [Citation44,Citation45] and others [Citation46].

In this work, we demonstrate a simple and controlled method for unzipping MWCNTs under continuous flow, depending on the operating conditions, in particular the rotational speed of the tube, choice of solvent, concentration of the MWCNTs, and the application of field effects for which the VFD is ideally designed. Herein this involved irradiating the VFD with a pulsed laser, using environmentally acceptable ethanol as the choice of solvent in which the MWCNTs were suspended. Moreover, the processing is under continuous flow, such that the scalability of the process is established as an integral part of optimizing the process from the outset, and the processing results in high yielding unzipping. The mechanism of unzipping the MWCNTs and the nature of the resulting material were examined using transmission electron microscopy (TEM), with the resulting nanocarbon material characterized using TEM, high-resolution transmission electron microscopy (HRTEM), scanning electron microscopy (SEM), thermogravimetric analysis (TGA), Raman spectroscopy, X-ray powder diffraction (XRD) and X-ray photoelectron spectroscopy (XPS).

2. Experimental section

2.1 Materials

MWCNTs were purchased from Sigma-Aldrich. The length and diameter of the as-received material were 3–6 µm and 10–50 nm, respectively. Ethanol with a purity of 99% was purchased from Chem-supply (HA 154-2).

2.2 Preparation of unzipping MWCNTs

The as-received MWCNTs (10 mg) were dispersed in ethanol (100 mL) over 30 min using bath sonication (6 KHz). A suspension of MWCNTs in ethanol was delivered via a stainless-steel jet feed close to the hemispherical bottom of the VFD tube rotating at high speed, while irradiated with a pulsed laser (Nd: YAG 1064 nm, Quanta Ray-LAB190, 8 mm diameter spot size) operating at 250 mJ as the optimal power (Figure ). The repetition rate and pulse width were 10 Hz and 5 ns, respectively. The beam of the laser was directed to the middle of VFD tube and the setup of the VFD and irradiation laser was according to a previously reported work on slicing CNTs under flow using the VFD [Citation39]. The VFD was tilted at 45° as the optimal angle for most processing in the device [Citation39,Citation40], with the concentration of the MWCNTs then systematically varied along with the rotational speed, which was initially set at 8k rpm, which is often optimal for VFD processing [Citation25,Citation34]. After unzipping MWCNTs, the yield was determined to be 90%, based on the volume and concentration of the starting materials of MWCNTs (0.45 mg mL−1) and the quantity of the isolated and dried unzipped MWCNTs.

Figure 1. Scheme illustrating the preparation of unzipping MWCNTs in a VFD irradiated with a pulsed laser, wavelength 1064 nm.

Figure 1. Scheme illustrating the preparation of unzipping MWCNTs in a VFD irradiated with a pulsed laser, wavelength 1064 nm.

2.3 Characterization

The morphology, size and properties of the unzipped MWCNTs were determined using SEM (FEI Quanta 450), TEM, HRTEM (Philips CM200), Raman spectroscopy (WiTec Alpha 300R λexc = 532 nm), TGA (Perkin Elmer STA8000). Thermogravimetric was carried out under nitrogen atmosphere from 50°C to 800°C at a rate of 10°C min−1, XPS (Phoibos 100 hemispherical analyser SPECS) with an Mg anode and XRD [Bruker Advanced D8 diffractometer (capillary stage)] using Co-Kα radiation.

3. Results and discussion

For laser-induced unzipping of MWCNTs in the VFD, the as-received materials were dispersed in ethanol at a concentration of 0.1 mg mL−1, and the solution then sonicated for 5 min in a bath sonicator prior to VFD processing. Then the solution was injected into the VFD using the continuous flow mode of the device, with the rotational speed at 8k rpm, tilt angle at 45° and injection rate of 0.45 mL min−1 Initially a pulsed laser with a power of 250 mJ and operating at a wavelength of 1064 nm was used to irradiate the VFD tube. A schematic illustrating the unzipping MWCNTs using laser irradiation with the VFD is presented in Figure .

Based on our previous findings [Citation34,Citation39], the shear stress within the VFD is effective for slicing CNTs in high yield with control over their length using an immiscible biphasic mixture of o-xylene and water [Citation26]. This was in the absence of laser irradiation, and the slicing is likely to involve contact electrification [Citation47,Citation48]. This is likely to prevail also in the most recent study in the VFD, establishing that shear stress acted as “scissors” to partially and completely unzip tubular structure of MWCNTs suspended in aqueous hydrogen peroxide (30% aqueous solution) which acts as an oxidizing agent. We sort to develop a one-step process for unzipping MWCNTs and producing GNRs in high yield using ethanol as a green solvent. In so doing several parameters, including MWCNT concentration and laser power, must be optimized during the VFD processing. Other operational parameters, such as the rotational speed (8k rpm), the tilt angle (45°) and the flow rate (0.45 mLmin−1), were found to be appropriate for processing carbon nanomaterials within the VFD [Citation39,Citation40,Citation49,Citation50]. Different concentrations of MWCNTs were studied, namely 0.5, 0.25 and 0.1 mg mL−1, as highlighted in the SEM images in Figures S1–S3. The concentration of 0.1 mg mL−1 was effective for complete unzipping of the MWCNTs, with the higher concentrations (0.25 and 0.5 mg mL−1) resulting in partial unzipping. It is important to note that processing materials with a high concentration in the VFD can disturb the high-shear topological fluid flow regimes which are responsible herein for the induced mechanical energy. This contributes to the unzipping on absorption of 1064 nm laser irradiation which serves to increase vibrational energy [Citation49,Citation51]. To optimize the laser power, four different laser powers of 100, 250, 400 and 600 mJ were explored. Pulsed laser irradiated at 250 mJ was the only power effective for completely unzipping the MWCNTs. Other laser powers did not dramatically change the tubular structure of MWCNTs, for both lower power, 100 mJ (Figure S2) and high power, 400 and 600 mJ (Figure S3).

The proposed mechanism for unzipping MWCNTs via laser irradiation with the VFD is illustrated as a scheme in Figure . The high-shear stress imparted in this film of liquid in the device and the effect of laser irradiation on the MWCNTs are two distinct processes, but they can act in concert. At high rotating speeds (8k rpm), the shear stress is dominated by the double-helical topological fluid flow in the VFD [Citation36]. At the double-helical flow, the VFD-generated mechanical energy can be used to slice CNTs and unzip MWCNTs depending on the choice of solvent(s). Interestingly, SWCNTs processed in the VFD can be coiled into regular toroidal structures which happen where the double helical flow strikes the surface of the tube, noting that the centrifugal force will accelerate the dense carbon material to the inner surface of the tube [Citation52]. The bending of the SWCNTs here provides stress on the C–C bonds. If at least bending of the MWCNTs used in the present study occurs, then this is possibly the starting point for the unzipping process in the presence of the induced vibrational energy from the pulsed laser. This is indeed the established mechanism for sling SWCNTs [Citation37]. Then why isn't there slicing of the MWCNTs in the present study? Presumably this is a trade-off between the ease of unzipping and exfoliating single shells of the MWCNTs versus slicing involving simultaneous C–C bond cleavage within all the centric rings in the material.

Figure 2. Proposed mechanism of laser-induced unzipping MWCNTs in the VFD in ethanol media using different laser powers: (a) 100 mJ, (b) 250 mJ (optimized conditions) and (c) 400 mJ.

Figure 2. Proposed mechanism of laser-induced unzipping MWCNTs in the VFD in ethanol media using different laser powers: (a) 100 mJ, (b) 250 mJ (optimized conditions) and (c) 400 mJ.

Low laser power (100 mJ) has little effect on the tubular structure of MWCNTs, as determined by TEM (Figure a). Increasing the pulsed laser irradiation to 250 mJ provides sufficient vibrational energy to result in C–C bond cleavage and unzipping. The heat generated by the irradiated laser plays a crucial role in unzipping MWCNTs, and when the VFD tube rapidly rotates, the aforementioned defects generated by the laser will expand and the unzipping process will be accelerated (Figure b). Increasing the laser power to 400 and 600 mJ results in more defects of MWCNT walls and the formation of irregular shapes, which can be attributed to the damaging of the graphitic structure and the formation of amorphous carbon, as shown in Figure (c). These results are in good agreement with our SEM and TEM results.

The unzipped MWCNTs were characterized using SEM, TEM, and HRTEM. The as-received MWCNTs exhibit entangled long tubes ranging in length from 3 to 6 µm and in diameter from 5 to 50 nm (Figure a–d). The tubular structure of as-received MWCNTs is evident from TEM and HRTEM images, and their outer diameter is ∼20 nm, as shown in Figure (c and d). The morphology of MWCNTs exposed to laser irradiation during VFD processing is dramatically changed, resulting in the formation of unzipped MMWCNTs. Figure (c–f) shows SEM images of the stated materials following VFD processing at 8k rpm using 250 mJ of laser irradiation. Low-magnification SEM images display unzipped MWCNTs with some GNRs and some small flat graphene sheets (Figure e and f). The low-magnification TEM image (Figure c) validates the unzipping process and displays MWCNTs that are completely unzipped and entangled. Figure (d and e) depict HRTEM images demonstrating that the width of unzipped MWCNTs is ∼30 nm. Figure (f) shows the HRTEM image of multitubes opened and formed unzipping MWCNTs.

Figure 3. SEM images with various magnifications of (a,b) as-received MWCNTs (before processing) and (c–f) GNRs (after processing) using the VFD at optimized conditions with laser power at 250 mJ.

Figure 3. SEM images with various magnifications of (a,b) as-received MWCNTs (before processing) and (c–f) GNRs (after processing) using the VFD at optimized conditions with laser power at 250 mJ.

Figure 4. TEM and HRTEM images of (a,b) as-received MWCNTs (before processing) and (c–f) unzipped MWCNTs (after processing) using laser irradiation (at 250 mJ) with the VFD operating at other optimized conditions.

Figure 4. TEM and HRTEM images of (a,b) as-received MWCNTs (before processing) and (c–f) unzipped MWCNTs (after processing) using laser irradiation (at 250 mJ) with the VFD operating at other optimized conditions.

The thermal stability of both materials was examined using TGA (Figure a). For the as received MWCNTs (black line), there is no weight loss until 600°C, followed by a significant weight loss between 600°C and 650°C, which is attributed to the decomposition of the MWCNTs [Citation53]. After processing in the VFD with laser irradiation and generating unzipped MWCNTs (red line), the weight loss begins at ∼200°C and decreases significantly due to the loss of oxygen functional groups [Citation54]. The unzipped MWCNTs should become oxidized as a result of the high laser irradiation during the VFD processing, as determined using Raman spectroscopy and XPS, as detailed below.

Figure 5. (a) TGA curves were carried out under nitrogen atmosphere from 50°C to 800°C at a rate of 10°C min−1, (b) Raman spectra and (c) XRD spectra of as-received MWCNTs (before processing) and unzipped MWCNTs (after processing) using laser irradiation (at 250 mJ) with the VFD operating at other optimized conditions.

Figure 5. (a) TGA curves were carried out under nitrogen atmosphere from 50°C to 800°C at a rate of 10°C min−1, (b) Raman spectra and (c) XRD spectra of as-received MWCNTs (before processing) and unzipped MWCNTs (after processing) using laser irradiation (at 250 mJ) with the VFD operating at other optimized conditions.

Raman spectroscopy was used to determine the defect level of the as received and the transformed unzipped MWCNTs. The Raman spectrum displays two bands (D band) at 1330 and (G band) 1576 cm−1, which are associated with disorders in the sp2 carbon network and C–C inside the graphite network, respectively [Citation55,Citation56]. To determine the fraction of graphite structure crystallites, we calculated the ratio between the D and G bands. After processing and unzipping the MWCNTs, the average ratio (ID/IG) increased from 1.12 in the as-received material (black line) to 1.50 on processing in the VFD. Results demonstrate that laser irradiation with high-shear stress from the device induces additional defects on the surface of MWCNTs, which results in completely unzipped MWCNTs [Citation53].

To further investigate the structure of unzipped MWCNTs, X-ray diffraction (XRD) patterns of the as received MWCNTs and unzipped MWCNTs at 250 mJ were conducted, as shown in Figure (c). For the as received MWCNTs, two main strong diffraction peaks at 30° and 51° can be assigned to the (002) and (100) planes, which reflects of the hexagonal graphite lattice [Citation27,Citation57]. Similar peaks can be seen after processing MWCNTs in the VFD at low laser power 100 mJ, indicating no changes to the MWCNTs structure (see Figure S7). However, irradiating at high laser power (400 mJ), Figure S7, results in significant changes to the structure, with no crystalline peaks from the XRD pattern and all peaks for MWCNTs disappeared and the presence of a broad peak at 25.3°, which is clearly confirmed that MWCNTs derived material is amorphous [Citation58], which is in agreement with TEM results. After the MWCNTs were unzipped using 250 mJ laser power “optimum power”, Figure c, two main sharp diffraction peaks are observed at ∼22° (001) and ∼34.8° (220), which indicate unzipping of the MWCNTs and the presence of defects [Citation27,Citation59–61]. Other peaks at 27.2° (331), 42.5° (111) and 47.2° (200) can be seen, which may be due to either more oxygen functional groups on the exfoliated layers or the presence of the catalyst on the unzipped MWCNT surface [Citation62]. Note that the graphitic (002) (100) peaks in the as received MWCNTs remains, albeit very weak, which indicates unzipping of the MWCNTs and the fabrication of GNRs [Citation3,Citation18]”.

The surface composition of both the as-received MWCNTs and the unzipped MWCNTs was analysed using XPS. XPS spectra of C 1s are shown in Figure . The C 1s spectrum was fitted with four major peaks, as follows: the peaks at ∼284.7 and ∼285.3 eV correspond to carbon C–C and C=C, respectively. The other two peaks C–O at ∼287 eV and C=O at ∼289.2 eV correspond to functional groups of oxygen [Citation63,Citation64]. After unzipping the MWCNTs involving laser-mediated VFD processing, the percentage of C–C decreases from 42.69% to 39.20%, while the percentage of C=C increased from 34.60% to 39.69%. This shows that more functional groups of oxygen and defects are formed on the unzipped MWCNTs during laser irradiation in the VFD (inserted table in Figure ). Additionally, the intensity of the C–O peak decreases slightly from 12.12 to 10.12, while the percentage of the C=C peak increases slightly from 5.91 to 6.73 after unzipping the MWCNTs. This can be attributed to an increase in oxygen functional groups during oxidation and unzipping when laser irradiated during the VFD processing.

Figure 6. XPS survey spectra and C 1s spectra for (a) as-received MWCNTs (before processing) and (b) unzipped MWCNTs (after processing) using the VFD at optimized conditions with the laser irradiation power at 250 mJ.

Figure 6. XPS survey spectra and C 1s spectra for (a) as-received MWCNTs (before processing) and (b) unzipped MWCNTs (after processing) using the VFD at optimized conditions with the laser irradiation power at 250 mJ.

4. Conclusion

We have established that pulsed laser irradiation at 1064 nm on the rapidly rotating quartz tube in a vortex fluidic device (VFD) can convert the tubule structure of MWCNTs in an ethanolic suspension into unzipped MWCNTs under continuous flow conditions. This process is convenient using low capital outlay and small footprint device technology, and it is scalable, achieving a high yield of 90%. Additionally, ethanol is used as a dispersion medium for MWCNTs, which can be considered as a green solvent as opposed to harsh chemicals that would contribute to a waste stream.

Supplemental material

Supplemental Material

Download MS Word (26 MB)

Acknowledgements

T.M.D.A extends his appreciation to Prof. Colin L. Raston from Flinders University in Australia for his valuable discussions and proofreading this work. T. M. D. A thanks Amjad Alotaibi for assistance with XPS measurements. T.M.D.A. thanks Dr. Ashley Slattery from Adelaide Microscopy for assistance with TEM and HRTEM imaging. The author also thanks Taibah University (Ministry of Education, Saudi Arabia) for help and support and Australian Research Council (DP200101105) and The Government of South Australia for support of this work.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Georgakilas V, Perman JA, Tucek J, et al. Broad family of carbon nanoallotropes: classification, chemistry, and applications of fullerenes, carbon dots, nanotubes, graphene, nanodiamonds, and combined superstructures. Chem Rev. 2015;115:4744–4822. doi:10.1021/cr500304f
  • Hong G, Diao S, Antaris AL, et al. Carbon nanomaterials for biological imaging and nanomedicinal therapy. Chem Rev. 2015;115:10816–10906. doi:10.1021/acs.chemrev.5b00008
  • Kosynkin DV, Higginbotham AL, Sinitskii A, et al. Longitudinal unzipping of carbon nanotubes to form graphene nanoribbons. Nature. 2009;458:872–876. doi:10.1038/nature07872
  • Bati AS, Yu L, Batmunkh M, et al. Synthesis, purification, properties and characterization of sorted single-walled carbon nanotubes. Nanoscale. 2018;10:22087–22139. doi:10.1039/C8NR07379A
  • Bati AS, Yu L, Batmunkh M, et al. Recent advances in applications of sorted single-walled carbon nanotubes. Adv Funct Mater. 2019;29:1902273. doi:10.1002/adfm.201902273
  • Kumar R, Singh RK, Singh DP, et al. Laser-assisted synthesis, reduction and micro-patterning of graphene: recent progress and applications. Coordination Chem Rev. 2017;342:34–79. doi:10.1016/j.ccr.2017.03.021
  • Kumar R, delPino AP, Sahoo S, et al. Laser processing of graphene and related materials for energy storage: state of the art and future prospects. Prog Energy Combust Sci. 2022;91:100981. doi:10.1016/j.pecs.2021.100981
  • Luo S, Chen X, He Y, et al. Recent advances on graphene nanoribbons for biosensing and biomedicine. J Mater Chem B. 2021;9:6129–6143. doi:10.1039/D1TB00871D
  • Xu W, Lee T-W. Recent progress in fabrication techniques of graphene nanoribbons. Mater Horiz. 2016;3:186–207. doi:10.1039/C5MH00288E
  • Terrones M. Sharpening the chemical scissors to unzip carbon nanotubes: crystalline graphene nanoribbons. ACS Nano. 2010;4:1775–1781. doi:10.1021/nn1006607
  • Kumar R, Joanni E, Savu R, et al. Fabrication and electrochemical evaluation of micro-supercapacitors prepared by direct laser writing on free-standing graphite oxide paper. Energy. 2019;179:676–684. doi:10.1016/j.energy.2019.05.032
  • Kumar R, Joanni E, Singh RK, et al. Direct laser writing of micro-supercapacitors on thick graphite oxide films and their electrochemical properties in different liquid inorganic electrolytes. J Colloid Interface Sci. 2017;507:271–278. doi:10.1016/j.jcis.2017.08.005
  • Lee KW, Lee CE. Electric field-induced unzipping of hydrogenated carbon nanotubes into graphene nanoribbons. Curr Appl Phys. 2014;14:337–339. doi:10.1016/j.cap.2013.12.015
  • Sinitskii A, Fursina AA, Kosynkin DV, et al. Electronic transport in monolayer graphene nanoribbons produced by chemical unzipping of carbon nanotubes. Appl Phys Lett. 2009;95:253108. doi:10.1063/1.3276912
  • Abbas AN, Liu G, Liu B, et al. Patterning, characterization, and chemical sensing applications of graphene nanoribbon arrays down to 5 nm using helium ion beam lithography. ACS Nano. 2014;8:1538–1546. doi:10.1021/nn405759v
  • Lee HJ, Lim J, Cho S-Y, et al. Intact crystalline semiconducting graphene nanoribbons from unzipping nitrogen-doped carbon nanotubes. ACS Appl Mater Interfaces. 2019;11:38006–38015. doi:10.1021/acsami.9b08876
  • Zheng Q-F, Guo Y, Liang Y, et al. Graphene nanoribbons from electrostatic-force-controlled electric unzipping of single-and multi-walled carbon nanotubes. ACS Appl Nano Mater. 2020;3:4708–4716. doi:10.1021/acsanm.0c00710
  • ZehtabYazdi A, Chizari K, Jalilov AS, et al. Helical and dendritic unzipping of carbon nanotubes: a route to nitrogen-doped graphene nanoribbons. ACS Nano. 2015;9:5833–5845. doi:10.1021/acsnano.5b02197
  • Xu X, Ruan S, Zhai J, et al. Facile synthesis of graphene nanoribbons from zeolite-templated ultra-small carbon nanotubes for lithium ion storage. J Mater Chem A. 2018;6:21327–21334. doi:10.1039/C8TA07501H
  • Kim K, Sussman A, Zettl A. Graphene nanoribbons obtained by electrically unwrapping carbon nanotubes. ACS Nano. 2010;4:1362–1366. doi:10.1021/nn901782g
  • Vadahanambi S, Jung J-H, Kumar R, et al. An ionic liquid-assisted method for splitting carbon nanotubes to produce graphene nano-ribbons by microwave radiation. Carbon. 2013;53:391–398. doi:10.1016/j.carbon.2012.11.029
  • Joanni E, Kumar R, Fernandes WP, et al. In situ growth of laser-induced graphene micro-patterns on arbitrary substrates. Nanoscale. 2022;14:8914–8918. doi:10.1039/D2NR01948E
  • Silva-Santos S, Alencar R, Aguiar A, et al. From high pressure radial collapse to graphene ribbon formation in triple-wall carbon nanotubes. Carbon. 2019;141:568–579. doi:10.1016/j.carbon.2018.09.076
  • Kumar P, Yamijala SS, Pati SK. Optical unzipping of carbon nanotubes in liquid Media. J Phys Chem. C. 2016;120:16985–16993. doi:10.1021/acs.jpcc.6b02524
  • Choi BG, Yang M, Hong WH, et al. 3D macroporous graphene frameworks for supercapacitors with high energy and power densities. ACS Nano. 2012;6:4020–4028. doi:10.1021/nn3003345
  • Silva AA, Pinheiro RA, Rodrigues AC, et al. Graphene sheets produced by carbon nanotubes unzipping and their performance as supercapacitor. Appl Surf Sci. 2018;446:201–208. doi:10.1016/j.apsusc.2018.01.214
  • Ko D, Choi J, Yan B, et al. A facile and scalable approach to develop electrochemical unzipping of multi-walled carbon nanotubes to graphene nanoribbons. J. Mater. Chem. A. 2020;8:22045–22053. doi:10.1039/D0TA03782F
  • Kumar P, Panchakarla L, Rao C. Laser-induced unzipping of carbon nanotubes to yield graphene nanoribbons. Nanoscale. 2011;3:2127–2129. doi:10.1039/c1nr10137d
  • Bang SY, Ryu JH, Choi BG, et al. A simple route for fabrication of graphene nanoribbons by pulsed laser irradiation in ethanol. J. Alloys Compd. 2014;618:33–36. doi:10.1016/j.jallcom.2014.08.004
  • Hamidinejad M, Zhao B, Zandieh A, et al. Enhanced electrical and electromagnetic interference shielding properties of polymer–graphene nanoplatelet composites fabricated via supercritical-fluid treatment and physical foaming. ACS Appl Mater Interfaces. 2018;10:30752–30761. doi:10.1021/acsami.8b10745
  • Arjmand M, Sadeghi S, OteroNavas I, et al. Carbon nanotube versus graphene nanoribbon: impact of nanofiller geometry on electromagnetic interference shielding of polyvinylidene fluoride nanocomposites. Polymers. 2019;11:1064. doi:10.3390/polym11061064
  • Jun Y-s, Habibpour S, Hamidinejad M, et al. Enhanced electrical and mechanical properties of graphene nano-ribbon/thermoplastic polyurethane composites. Carbon. 2021;174:305–316. doi:10.1016/j.carbon.2020.12.023
  • Lin G, Zhou T, Zhou Z, et al. Laser induced graphene for EMI shielding and ballistic impact damage detection in basalt fiber reinforced composites. Compos Sci Technol. 2023;242:110182.
  • Alharbi TM, Alotaibi AE, Chen D, et al. Unzipping multiwalled carbon nanotubes under vortex fluidic continuous flow. ACS Appl Nano Mater. 2022;5:12165–12173. doi:10.1021/acsanm.2c02448
  • Luo X, Al-Antaki AHM, Vimalanathan K, et al. Laser irradiated vortex fluidic mediated synthesis of luminescent carbon nanodots under continuous flow. React Chem Eng. 2018;3:164–170. doi:10.1039/C7RE00197E
  • Alharbi T, Jellicoe M, Luo X, et al. Sub-micron moulding topological mass transport regimes in angled vortex fluidic flow. Nanoscale Adv. 2021;3:3064–3075. doi:10.1039/D1NA00195G
  • Vimalanathan K, Raston CL. Dynamic thin films in controlling the fabrication of nanocarbon and its composites. Adv Mater Technol. 2017;2:1600298. doi:10.1002/admt.201600298
  • Britton J, Stubbs KA, Weiss GA, et al. Vortex fluidic chemical transformations. Chem Eur J. 2017;23:13270–13278. doi:10.1002/chem.201700888
  • Alharbi TM, Li Q, Raston CL. Thin film mechano-energy induced slicing of carbon nanotubes under flow. ACS Sustain Chem Eng. 2021;9:16044–16051. doi:10.1021/acssuschemeng.1c03109
  • Alharbi TM, Vimalanathan K, Lawrance WD, et al. Controlled slicing of single walled carbon nanotubes under continuous flow. Carbon. 2018;140:328–434. doi:10.1016/j.carbon.2018.08.066
  • Batmunkh M, Vimalanathan K, Wu C, et al. Efficient production of phosphorene nanosheets via shear stress mediated exfoliation for low-temperature perovskite solar cells. Small Methods. 2019;3:1800521. doi:10.1002/smtd.201800521
  • Vimalanathan K, Suarez-Martinez I, Peiris C, et al. Vortex fluidic mediated transformation of graphite into highly conducting graphene scrolls. Nanoscale Adv. 2019;1:2495–2501. doi:10.1039/C9NA00184K
  • Al-Antaki AHM, Luo X, Alharbi TM, et al. Inverted vortex fluidic exfoliation and scrolling of hexagonal-boron nitride. RSC Adv. 2019;9:22074–22079. doi:10.1039/C9RA03970H
  • Alharbi TM, Shingaya Y, Vimalanathan K, et al. High yielding fabrication of magnetically responsive coiled single walled carbon nanotube under flow. ACS Appl Nano Mater. 2019;2:5282–5289. doi:10.1021/acsanm.9b01135
  • Alsulam IK, Alharbi TM, Moussa M, et al. High-yield continuous-flow synthesis of spheroidal C60@ graphene composites as supercapacitors. ACS Omega. 2019;4:19279–19286. doi:10.1021/acsomega.9b02656
  • Yasmin L, Chen X, Stubbs KA, et al. Optimising a vortex fluidic device for controlling chemical reactivity and selectivity. Sci Rep. 2013;3:2282. doi:10.1038/srep02282
  • Lin S, Chen X, Wang ZL. Contact electrification at the liquid–solid interface. Chem Rev. 2022;112:5209–5232. doi:10.1021/acs.chemrev.1c00176
  • Kaponig M, Mölleken A, Nienhaus H, et al. Dynamics of contact electrification. Sci Adv. 2021;7:595. doi:10.1126/sciadv.abg7595
  • Vimalanathan K, Gascooke JR, Suarez-Martinez I, et al. Fluid dynamic lateral slicing of high tensile strength carbon nanotubes. Sci Rep. 2016;6:22865. doi:10.1038/srep22865
  • Alharbi TM, Harvey D, Alsulami IK, et al. Shear stress mediated scrolling of graphene oxide. Carbon. 2018;137:419–424. doi:10.1016/j.carbon.2018.05.040
  • Alharbi T, Vimalanathan K, Alsulami I, et al. Vertically aligned laser sliced MWCNTs. Nanoscale. 2019;11:21394–21403.
  • Jellicoe M, Gibson CT, Quinton JS, et al. Coiling of single-walled carbon nanotubes via selective topological fluid flow: implications for sensors. ACS Appl Nano Mater. 2022;5:11586–11594. doi:10.1021/acsanm.2c02579
  • Xiao B, Li X, Li X, et al. Graphene nanoribbons derived from the unzipping of carbon nanotubes: controlled synthesis and superior lithium storage performance. J Phys Chem C. 2014;118:881–890. doi:10.1021/jp410812v
  • Dimiev AM, Khannanov A, Vakhitov I, et al. Revisiting the mechanism of oxidative unzipping of multiwall carbon nanotubes to graphene nanoribbons. ACS Nano. 2018;12:3985–3993. doi:10.1021/acsnano.8b01617
  • Yang M, Hu L, Tang X, et al. Longitudinal splitting versus sequential unzipping of thick-walled carbon nanotubes: towards controllable synthesis of high-quality graphitic nanoribbons. Carbon. 2016;110:480–489. doi:10.1016/j.carbon.2016.09.055
  • Wang C, Li Y-S, Jiang J, et al. Controllable tailoring graphene nanoribbons with tunable surface functionalities: an effective strategy toward high-performance lithium-ion batteries. ACS Appl Mater interfaces. 2015;7:17441–17449. doi:10.1021/acsami.5b04864
  • Liu M, Song Y, He S, et al. Nitrogen-doped graphene nanoribbons as efficient metal-free electrocatalysts for oxygen reduction. ACS Appl Mater Interfaces. 2014;6:4214–4222. doi:10.1021/am405900r
  • Sivakumar A, JudeDhas SS, Pazhanivel T, et al. Phase transformation of amorphous to crystalline of multiwall carbon nanotubes by shock waves. Cryst Growth Des. 2021;21:1617–1624. doi:10.1021/acs.cgd.0c01464
  • Romero Aburto R, Alemany LB, Weldeghiorghis TK, et al. Chemical makeup and hydrophilic behavior of graphene oxide nanoribbons after low-temperature fluorination. ACS Nano. 2015;9:7009–7018. doi:10.1021/acsnano.5b01330
  • Li J, Ye S, Li T, et al. Preparation of graphene nanoribbons (GNRs) as an electronic component with the multi-walled carbon nanotubes (MWCNTs). Procedia Eng. 2015;102:492–498. doi:10.1016/j.proeng.2015.01.197
  • Medina P, Zheng H, Fahlman B, et al. Li4Ti5O12/graphene nanoribbons composite as anodes for lithium ion batteries. SpringerPlus. 2015;4:1–7. doi:10.1186/s40064-015-1438-0
  • He X, Xu X, Bo G, et al. Studies on the effects of different multiwalled carbon nanotube functionalization techniques on the properties of bio-based hybrid non-isocyanate polyurethane. RSC Adv. 2020;10:2180–2190. doi:10.1039/C9RA08695A
  • Estrade-Szwarckopf H. XPS photoemission in carbonaceous materials: A “defect” peak beside the graphitic asymmetric peak. Carbon. 2004;42:1713–1721. doi:10.1016/j.carbon.2004.03.005
  • Shu Q, Xia Z, Wei W, et al. Controllable unzipping of carbon nanotubes as advanced Pt catalyst supports for oxygen reduction. ACS Appl Energy Mater. 2019;2:5446–5455. doi:10.1021/acsaem.9b00506