量化石墨烯纳米带的排列和质量:偏振拉曼光谱方法

时间:2024-03-02 14:26:45 浏览量:0

ABSTRACT  

Graphene nanoribbons (GNRs) are atomically precise stripes of graphene with tunable electronic properties,  making them promising for room-temperature switching applications like field-effect transistors (FETs). However, challenges persist in GNR processing and characterization, particularly regarding GNR alignment during  device integration. In this study, we quantitatively assess the alignment and quality of 9-atom-wide armchair  graphene nanoribbons (9-AGNRs) on different substrates using polarized Raman spectroscopy. Our approach  incorporates an extended model that describes GNR alignment through a Gaussian distribution of angles. We not  only extract the angular distribution of GNRs but also analyze polarization-independent intensity contributions  to the Raman signal, providing insights into surface disorder on the growth substrate and after substrate transfer.  Our findings reveal that low-coverage samples grown on Au(788) exhibit superior uniaxial alignment compared  to high-coverage samples, attributed to preferential growth along step edges, as confirmed by scanning tunneling  microscopy (STM). Upon substrate transfer, the alignment of low-coverage samples deteriorates, accompanied by  increased surface disorder. For high-coverage samples, the alignment is preserved, and the disorder on the target  substrate is reduced compared to the low-coverage samples. Our extended model enables a quantitative  description of GNR alignment and quality, facilitating the development of GNR-based nanoelectronic devices.


1. Introduction 

 Graphene nanoribbons (GNRs) are quasi-one-dimensional stripes of  graphene with an intriguing set of physicochemical properties deriving  from quantum confinement and related bandgap tunability. The  ability to tune the properties of GNRs at the atomic scale by changing  their width and edge structure has opened up a promising  avenue for their application in electronics, spintronics,  and photonics. The required atomic precision in GNR synthesis  could only be met by a bottom-up approach based on the covalent  coupling of specifically designed precursor molecules followed by  cyclodehydrogenation on metallic surfaces. Since the pioneering work  of Cai et al. in 2010 , GNRs with various widths, edge  topologies (armchair , zigzag, cove , etc.), as well as specific edge extensions giving rise to exotic topological quantum phases, have been reported.


To explore the exciting properties of GNRs in functional devices, a  substrate transfer step is necessary to transfer the GNRs from their  metallic growth substrate (usually Au(111)) to semiconducting or  insulating substrates suitable for digital logic applications, such as SiO2/  Si . Most of the substrate transfer strategies developed so far  involve aqueous solutions or the presence of polymers as a support layer,  which can lead to residues or defects in the GNRs . To successfully  integrate GNRs into devices, GNR properties must be preserved and  monitored, also upon substrate transfer, which remains one of the main  bottlenecks in the development of GNR-based electronics.


2. Methods  

2.1. On-surface synthesis and STM characterization of 9-AGNRs  

The Au(788) single crystal growth substrate (MaTecK GmbH, Germany) was cleaned in ultra-high vacuum (UHV) with two cycles of  sputtering at1 kV Ar+ for 10 min and annealing at 420 ◦C for 10 min. The  9-AGNR precursor monomer 3′,6′-di-iodine-1,1′:2′,1″-terphenyl (DITP)  was then sublimated onto the clean Au surface from a quartz crucible heated to 70 ◦C while the substrate remained at room temperature.  A quartz microbalance was used to control the deposition rate of the  precursor molecules at 1 Å/min. The deposition rate is not calibrated to  correspond to the true surface coverage, but only give a relative measurement that is then calibrated by STM. High- and low-coverage samples were obtained by DITP deposition for 8 and 3 min, respectively.  Following deposition, the substrate was heated to 200 ◦C (0.5 K/s) for  10 min to initiate DITP polymerization, followed by annealing at 400 ◦C  (0.5 K/s) for 10 min to form the GNRs by cyclodehydrogenation.


2.2. Raman spectroscopy  

Raman spectroscopy measurements were obtained using a WITec  confocal Raman microscope (WITec Alpha 300R) with a laser line of  785 nm (1.5 eV) and a power of 40 mW. A 50 × microscope objective  (0.55 numerical aperture) with a working distance of 9.1 mm was used  to focus the laser beam onto the sample and collect the scattered light.  Calibration of Raman spectra was performed using the Si peak at 520.5  cm−1 . Also, the laser wavelength, power, and integration time were  optimized for each substrate to maximize signal while minimizing  sample damage. Furthermore, to avoid sample damage, a Raman mapping approach with 10 × 10 pixels (10 × 10 μm) was used and samples  were measured in a home-built vacuum suitcase with pressure ~10− 2  mbar. The vacuum chamber was mounted on a piezo stage for scanning.  


3. Results and discussion  

To synthesize aligned 9-AGNRs the precursor monomer 3′,6′-diiodine-1,1′:2′,1″-terphenyl (DITP)  is deposited on a vicinal catalytic  surface (Au(788)) followed by two annealing steps to activate the  polymerization and cyclodehydrogenation reactions. Samples  are prepared with two different coverages (~0.4 of a monolayer and ~1  full monolayer, ML, herein referred to as low- and high-coverage samples, respectively) as shown in Fig. 1. The vicinal surface enables the  growth of GNRs along the low-coordination sites of the Au(788) step  edges, which act as favorable nucleation sites. This allows GNRs  to grow gradually with deposition time, and after 8 min with a fixed  deposition rate of 1 Å/min, a full monolayer of aligned 9-AGNRs  (high-coverage sample) is formed. A representative STM image of a  high-coverage sample with 9-AGNRs of an average length of 34 nm is  shown in Fig. 1a, the corresponding GNR length histogram is given in  Fig. S1a. For the low-coverage 9-AGNR samples, a deposition time of 3  min is used (with a fixed deposition rate of 1 Å/min), which provides  just enough precursor molecules for individual 9-AGNRs to grow along  all Au(788) step edges, resulting in an average GNR length of 37 nm  (Fig. 1c, see Fig. S1b for the length histogram).


图片1

Fig. 1. Characterization of aligned 9-AGNRs samples at high (a,b) and low (c,d) coverage before (a,c, blue spectra) and after (b,d, red spectra) substrate transfer.  Raman spectra of the high-coverage sample on Au(788) (a) and after substrate transfer onto a Raman-optimized substrate (ROS)(b). The spectra are acquired  with an excitation wavelength of 785 nm under vacuum conditions with polarization parallel (I‖) to the GNR alignment direction (along the Au(788) step edges) (full  line), and perpendicular (I⊥) to the GNR alignment direction (dashed line). The inset in panel (a) shows a STM topography image for the high-coverage sample on Au  (788) (black arrow highlighting the GNR growth direction and position) with a scale bar of 10 nm (Vb = -1.5 V, It = 0.3 nA). The inset in panel (b) shows an optical  micrograph of ribbons transferred onto a ROS, with a scale bar of 180 μm. Raman spectra of the low-coverage sample (c) on Au(788), and (d) after substrate transfer  onto a ROS, with polarizations/full vs dashed lines as indicated above. The inset in panel (c) shows a STM image for the low-coverage sample on Au(788), with a  black arrow highlighting the GNR growth direction and position along the Au(788) step edges of (Vb = -1.5 V, It = 0.3 nA, scale bar: 10 nm). The inset in panel (d)  shows an optical micrograph of GNRs transferred onto a ROS, with a scale bar of 180 μm. All Raman spectra show four main modes: RBLM (width-dependent mode),  CH (C–H bending mode at the edges), D (breathing mode of the sp2 lattice) and G (stretching of C–C bonds within the sp2 lattice).


Raman profiles are obtained by polarizing the incoming and scattered light in parallel (“VV configuration”) with different angles between the nominal GNR alignment direction and the polarization of the  incident light. Using the VV configuration implies that for the Raman  resonant modes, the intensity of the GNR modes is projected to be  cos4 (ϑ) polarization-dependent, which results from a product of two  cos2 (ϑ) factors, one for photon absorption and the other for photon  emission, Eq. (1) . This means that the Raman signal is  maximum with the incident polarization parallel to the ribbon axis (0◦,  180◦) and zero when perpendicular to it (90◦, 270◦) (Fig. 2). In Eq. (1),  ϑ0 is the orientation of the long axis of the GNR with respect to an  arbitrary in-plane axis, and ϑ is the direction of the light polarization.  Due to the significant absorption anisotropy of the quasi-1D GNRs, all  Raman modes exhibit roughly the same polarization dependency.


图片2

Fig. 2. Illustration of the major contributions to the expected polarized Raman  intensity as described by Eq.. The Raman intensity of a single GNR follows  cos4 (ϑ) dependence (red). In blue, the angular distribution function D(ϑ) is  shown, which includes the normalized Gaussian distribution of angles and the  normalized isotropic contribution. For this particular plot, σ = 3◦, A = 0.5, and  B = 0.5 have been used.  


To investigate the influence of GNR coverage and substrate transfer  on σ and OD, we fit all Raman active modes of high- and low-coverage 9-  AGNR samples on both the growth and ROS using Eq. (9). Fig. 3 shows  the G mode peak intensity as a function of the polarization angle ϑ for  the VV configuration and the related polar diagrams for both high-  (Fig. 3a–b) and low-coverage samples (Fig. 3c–d) on Au(788) (in blue)  and after substrate transfer (in red), respectively (see Figs. S4 and S5 for  similar plots for CH, D, and RBLM modes). The intensity of the G mode  as a function of polarization angle (− 90◦ to +90◦) is determined from  Raman maps of 10 x 10 pixels in vacuum conditions using a 785 nm laser  energy.


图片3

Fig. 3. Polarized Raman intensity of G mode (785 nm, VV configuration). (a, c) G mode intensity as a function of polarization angle ϑ for high- and low-coverage  samples on Au(788) (blue circles) and after substrate transfer onto ROS (red squares). Blue and red solid lines represent data fits using Eq. (9). (b, d) Polar diagrams  showing G mode intensities for high- and low-coverage samples on Au(788) (blue circles) and after transfer to ROS (red squares). Blue and red solid lines represent  fits to the measured data using Eq. (9).


On the Au(788) growth substrate we observe a similar OD for both  high- and low-coverage samples (8 %). The disorder observed on the  growth substrate may originate from short (and thus non-aligned) GNRs,  irregularly fused precursor monomers, or also the presence of impurities  from the precursor monomer.


4. Conclusions  

In this study, we employed polarized Raman spectroscopy and  scanning tunneling microscopy to characterize and quantify the structural quality and degree of alignment of 9-AGNRs in samples with  different surface coverages on both their growth substrate and after  substrate transfer. Using an extended data analysis model, which describes GNR alignment by a Gaussian distribution of angles, allowed us  to extract both the quality of alignment (σ) and the overall surface disorder (OD).


Our results show that low-coverage samples exhibit better uniaxial  alignment than high-coverage samples on the growth substrate. This  behavior results from GNRs in low-coverage samples growing preferentially along the step-edges of Au(788), as observed in our STM investigations. However, upon transfer, the quality of alignment of lowcoverage samples is significantly reduced, which we attribute mostly  to the strong interaction of GNRs with the Au(788) step edges as well as  increased GNR mobility, whereas high-coverage samples show better  alignment preservation upon substrate transfer, owing to the densely  packed GNR film facilitating the transfer process. With the extended model developed in this study, we also quantified the OD, which results  in an isotropic (polarization-independent) contribution to the Raman  intensity. After substrate transfer, low-coverage samples show systematically higher OD values than high-coverage samples (39 % vs 13 %  respectively). The significantly higher OD for low-coverage samples is  associated with the strong interaction of GNRs to the Au(788) step  edges, making it less likely for the GNRs to transfer efficiently, as well as  to the fact that more gold surface area is exposed to PMMA and other  impurities that may react with the metal and transfer along with the  GNRs to the target substrate. Based on these findings, strategies to  improve GNR alignment and quality are needed. One approach could be  the passivation of Au(788) step edges with other materials, such as wide  bandgap polymers. The presence of a polymer at the step edges could  simultaneously decrease the strong interaction between GNR-Au and act  as a scaffold, mitigating GNR’s lateral diffusion and preserving GNR  alignment throughout the substrate transfer process.

文件下载请联系管理员: 400-876-8096