Next Article in Journal
Improving Plasmonic Photothermal Therapy of Lung Cancer Cells with Anti-EGFR Targeted Gold Nanorods
Next Article in Special Issue
Gas Sensors and Semiconductor Nanotechnology
Previous Article in Journal
Direct Patterned Zinc-Tin-Oxide for Solution-Processed Thin-Film Transistors and Complementary Inverter through Electrohydrodynamic Jet Printing
Previous Article in Special Issue
NH3-Sensing Mechanism Using Surface Acoustic Wave Sensor with AlO(OH) Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Orientation Ordering and Chiral Superstructures in Fullerene Monolayer on Cd (0001)

School of Physical Science and Technology, Southwest University, Chongqing 400715, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this paper.
Nanomaterials 2020, 10(7), 1305; https://doi.org/10.3390/nano10071305
Submission received: 27 May 2020 / Revised: 21 June 2020 / Accepted: 30 June 2020 / Published: 3 July 2020
(This article belongs to the Special Issue Gas Sensors and Semiconductor Nanotechnology)

Abstract

:
The structure of C60 thin films grown on Cd (0001) surface has been investigated from submonolayer to second monolayer regimes with a low-temperature scanning tunneling microscopy (STM). There are different C60 domains with various misorientation angles relative to the lattice directions of Cd (0001). In the (2√3 × 2√3) R30° domain, orientational disorder of the individual C60 molecules with either pentagon, hexagon, or 6:6 bond facing up has been observed. However, orientation ordering appeared in the R26° domain such that all the C60 molecules adopt the same orientation with the 6:6 bond facing up. In particular, complex chiral motifs composed of seven C60 molecules with clockwise or anticlockwise handedness have been observed in the R4° and R8° domains, respectively. Scanning tunneling spectroscopy (STS) measurements reveal a reduced HOMO–LOMO gap of 2.1 eV for the C60 molecules adsorbed on Cd (0001) due to the substrate screening and charge transfer from Cd to C60 molecules.

1. Introduction

As a prototypical fullerene molecule, C60 can be used as the building blocks for carbon-based nanomaterials. A variety of C60 monolayers grown on solid surfaces is critical for understanding and controlling the interface properties of fullerene-derived electronic and photovoltaic devices [1,2,3]. In the past decades, there have been a large number of investigations of C60 monolayer structures grown on a wide range of metallic or semiconducting substrates such as Ag [4,5,6,7,8,9,10,11], Au [12,13,14,15,16,17,18,19,20,21,22,23], Cu [24,25,26,27,28], graphene [29,30,31,32,33], Si [34,35,36], Ge [37,38,39], or NaCl [40]. It was found that almost all monolayers of C60 adopt a close-packed structure regardless whether the substrate is isotropic or anisotropic, metallic or nonmetallic.
The C60 domains can take different orientations with respect to the directions of substrate lattices. On noble metal surfaces such as Au(111) or Ag(111), STM studies have demonstrated that the C60 monolayer domains exhibit a variety of lattice orientations such as the “in phase” (7 × 7) R0° [21], (2√3 × 2√3) R30° [4,8,9,10,18,19,20,21,22], and (√589 × √589) R14.5° [12,14,21] phases, which contain 4, 1, and 49 molecules per unit cell, respectively. In particular, “bright” and “dim” molecules have been found in the (2√3 × 2√3) R30° commensurate domains. It was proposed that the dim molecules are located on single atom vacancies, while the bright molecules remain on top position of the unreconstructed Au (111) surface. Moreover, a uniform R30° domain was also observed for all C60 molecules with the same contrast [15,21].
In addition to the different close-packing directions, the orientations of individual molecules within a single domain can also be different. The individual C60 molecules within a single domain can display multiple orientations such that the orientational ordering can be observed at 78 K [35]. For example, a complex orientational ordering was observed for molecules inside the “in-phase” (R0°) domain after room-temperature deposition, where a 7-molecule cluster is composed of a central molecule and six tilted surrounding molecules [18]. In the R14.5° domain, there are 49 molecules with a number of different orientations, on average, every time the seventh C60 appeared dim, giving rise to a quasiperiodic 7 × 7 superstructure at 5.7 K [12]. In particular, a chiral superstructure made of 7-molecule pinwheels was identified in C60 multilayer film on NaCl at 77 K [40] and in KxC60 monolayer on Au (111) at 7 K [41]. The proposed mechanism for orientational ordering was attributed to the maximized overlapping of neighboring molecular orbitals driven by the superexchange interaction. The latter refers to virtual hopping of an electron from the HOMO of one C 60 4 to the LUMO of its nearest neighbor to gain an energy proportional to the hopping amplitude and the inverse of HOMO–LUMO gap.
In this paper, we chose the divalent metal Cd (0001) thin films as the substrate. The hexagonal close-packed metal Cd is usually used as an electrode material due to the smaller electronegativity compared with precious metals [42,43,44]. Thus, it would be fundamentally important to study the interfaces structure of C60 monolayer on Cd (0001) because the C60 molecules need to be contacted with metallic electrode in the electronic devices. It is found that the C60 domains exhibit a variety of orientations with respect to the lattice directions of Cd (0001). In the (2√3 × 2√3) R30° domain, individual C60 molecules reveal orientational disorder with the pentagon, hexagon, or 6:6 bond facing up. In the R26° domain, orientation ordering takes place such that all the C60 molecules adopt the same orientation with the 6:6 bond facing up. More interestingly, the complex chiral motifs composed of seven C60 molecules with clockwise or anticlockwise handedness have been observed in the R4° and R8° domains. STS measurements shows that the C60 molecules have a reduced HOMO–LOMO gap of 2.1 eV, indicating a significant charge transfer from Cd to C60 and strong substrate screening effect.

2. Experiment

The experiments were conducted in an ultra-high vacuum, low-temperature scanning tunneling microscopy (Unisoku USM1500) with a base pressure 1.2 × 10−10 mbar. The Cd (0001) thin films were grown on the Si (111)-7 × 7 surface at room temperature. The Si (111) substrate was continuously degassed at ~870 K for 8 h with subsequent flashing to 1400 K for several seconds. Cd atoms with a purity of 99.998% were thermally evaporated onto the Si (111) substrate from a quartz crucible by controlling the current. Deposition of 15 monolayer (ML) Cd at room temperature results in flat and smooth Cd (0001) films, which exhibit a perfect transparency due to the strong anisotropic electron motion with large lateral effective mass [45]. C60 were evaporated from a homemade tantalum boat at a rate of 0.4 ML/min with the temperature of 770 K. After the deposition, the sample was transferred into the LT-STM chamber. All the STM images were obtained in constant-current mode at 78 K. All differential conductance dI/dV spectra were acquired using a standard lock-in amplifier with a bias modulation signal of 10 mV at 1999 Hz under open-loop condition.

3. Results and Discussion

We firstly studied the initial adsorption stage of C60 molecules on Cd (0001). Figure 1a shows the STM image of an isolated C60 molecule, which has a faint nodal plane separating the round protrusion into two lobes. The insert in Figure 1a shows the atomic-resolution image of the hexagonal lattices of Cd (0001) thin films grown on Si(111)–7 × 7 The measured lattice constant of Cd (0001) films is 3.0 ± 0.05 Å, very close to the bulk value (2.973 Å) in Cd crystals [44]. Figure 1b shows the nucleation of C60 molecules around a 2D Cd island. It was found that the C60 molecules aggregated at the step edge of the Cd island, due to the increased interaction with substrate at these positions. From the zoomed-in image in Figure 1c, anisotropy can be found from the individual C60 molecules attaching to the Cd island: most molecules show an anisotropic shape due to the tip convolution effect in the fast-scanning direction, while keep the original length (~1 nm) parallel to the step edge. In fact, such a type of step decoration was previously observed in the C60 nanostructures grown on metal surfaces. The C60 molecules initially adsorbed at intersections of multiple steps and edges of monatomic steps on narrow terraces and periodic arrays of short chains formed as the coverage is increased [46]. When C60 molecules attach on Ag (111) island, a complete, single-strand C60 “necklace” circling the island formed. The decoration, in turn, made the equilibrium island shape round [47]. Shown in Figure 1d is a monolayer island of C60 with a close-packed hexagonal structure. There is a misorientation angle of 20° between the lattice directions of C60 island and Cd (0001) surface. The error limit in the measurement of orientation angle is ±2° in our experiments. The measured lattice constant is a 1 = 10.2 ± 0.1     , larger than the lattice constant (10.02 Å) of the (111) plane in fcc C60 crystals [48]. The apparent height of C60 molecules in this island is measured to be 7.0 ± 0.1 Å, at the bias of 2.2 V.
When the coverage increased to one monolayer, C60 molecules form a close-packed (2√3 × 2√3) R30° domain as shown in Figure 2a. The apparent height of C60 molecules in this domain is measured to be 9.0 ± 0.1 Å, at the bias of 2.2 V. The measured intermolecular spacing is increased to 10.4 ± 0.1 Å, much larger than the preferred spacing of in C60 crystals [48]. It implies that all the C60 molecules in the (2√3 × 2√3) R30° domain suffer a tensile strain as large as 3.8%, which would make the R30° domain less stable. Moreover, we noticed that all the C60 molecules present a uniform height except a few dim molecules. The height difference between the dim and normal C60 molecules is ~0.6 Å. In particular, some dim molecules aggregate into a wire-like region, indicating the attractive force among the dim molecules. Figure 2b is the high-resolution STM image of the uniform R30° domain, where C60 molecules with different molecular orientations can be identified. The individual molecules marked by red, green, and yellow circles exhibit a round protrusion with a small hole (like a doughnut), three lobes like a clover, and two parallel lobes corresponding to the C60 molecules with a pentagon, hexagon, and 6:6 bond facing up, respectively [13,40]. We noticed that such types of orientational disorder also appeared in the R30° domains of C60 grown on Au (111) [13,18,21,22,23] or Ge (111) surfaces [39].
When the lattice direction of C60 monolayer deviates from the Cd (0001) lattice for 26° angle, all C60 molecules adopt the same orientation, forming a homogeneous orientation domain. It is observed that from the empty-state STM image at 1.4 V (Figure 3a), all the C60 molecules present a two-lobe structure oriented at the same direction, similar to the motifs marked by yellow circle in Figure 2b. It means that all the molecules in R26° domain adopt the 6:6 bond facing-up orientation, and are adsorbed at the equivalent sites of the Cd (0001). The apparent height of C60 molecules in this domain is measured to be 8.7 ± 0.1 Å at the bias of −2.8 V. The intermolecular spacing is 10 ± 0.2 Å, slightly smaller than that in the C60 island of Figure 1d. Although a similar phenomenon was observed previously in the close-packed C60 monolayer grown on Au (111) [21], the bias-dependent behaviors of the homogeneous domain has not been studied yet. Here, we have made investigations of the variation of C60 molecular contrast with the bias voltage from 1.4 to −2.0 V in the R26° domain. As shown in Figure 3b, when the bias is reduced to 0.6 V, the contrast of each molecule shows four small protrusions: three bright and one dim. In the filled state STM image at −1.4 V (Figure 3c), all the C60 molecules present a bright protrusion but with a small off-center hole, resembling the contrast of C60 molecule with a carbon atom at the top [35,49]. When the bias voltage is increased to −2.0 V, the off-center hole becomes smaller and the bright protrusion becomes larger (Figure 3d). It is well known that STM image yields a spatial map of the local density of states (DOS) for the adsorbates on metal surfaces. Thus, the bias-dependent STM images mentioned above can be attributed to the DOS variation of C60 molecules with energy. However, the possibility of orientation modification driven by the tip electric field cannot be safely excluded under the successive STM scanning. Previously, the tip-induced rotation was reported in the unstable C60 molecules adsorbed on graphene/Cu (111), where tip electron field or tunneling electron might play an important role for the reorientation [32].
Beside the homogeneous orientation, we also found a local and complex orientational ordering in the C60 monolayer domain. As shown in Figure 4a, four chiral motifs composed of 7 C60 molecules appeared in the R (−4°) domain. The central C60 molecule presents three lobe protrusions, the other six tilted surrounding molecules appear as two-lobe motifs with a nodal plane aligned at different directions, constituting a pinwheel pattern with clockwise handedness. Figure 4b shows a C60 domain with a misorientation angle of +8° relative to the lattice directions of Cd (0001). There are eleven chiral pinwheels with anticlockwise handedness in this domain. Such types of 7-molecule clusters resemble the C60 clusters appeared in the R0° domain grown on Au (111) [18], but the chiral feature appeared in the present domain. To some extent, it is more like the chiral motifs appeared in the C60 multilayer on NaCl [40] or K4+δ C60 monolayer on Au (111) surface [41], which was attributed to overlap of neighboring molecular orbitals due to the superexchange interaction. The latter refers to virtual hopping of an electron from the HOMO of one C 60 4 to the LUMO of its nearest neighbor to gain an energy proportional to the hopping amplitude and the inverse of HOMO–LUMO gap. Both the apparent heights of C60 molecules in these two domains are 9.5 ± 0.1 Å. Furthermore, we found the C60 molecules in above two domains reveal the same intermolecular distance as that in Figure 3, i.e., 10 ± 0.2 Å.
Figure 5 shows a scanning tunneling spectra (STS) acquired on top of a C60 molecule at 77 K in the (2√3 × 2√3) R30° domain. The measured highest occupied molecular orbital (HOMO) is located at −1.8 eV, while the lowest unoccupied molecular orbital (LUMO) appears at 0.32 eV. Thus the HOMO–LUMO gap is 2.12 eV, which is much smaller than that (4.9 eV) of a free C60 molecule in gas state, but comparable to that (2.7 eV) of C60 on Au (111) and (2.3 eV) of C60 on Ag (100) [5,13]. Compared with the LUMOs of C60 on noble metal surfaces [0.84 eV for Au (111), 0.49 eV for Ag (100), and 0.43 eV for Cu (100)], the lower position of C60 LUMO (0.32 eV) on Cd (0001) reflect the stronger reactivity of Cd substrate [4,5,13]. Moreover, the effect of substrate screening may also reduce the HOMO–LUMO gap, through reducing the intramolecular Coulomb repulsion [50]. In addition to the peaks corresponding to HOMO and LUMO, there are also small peaks appeared in the gap, which result from the surface states or quantum well state in the Cd (0001) films [45].
We have also studied the structure of the second C60 layer. Shown in Figure 6a is a C60 trimer that appeared on top of the first monolayer. It was found that all the three molecules in the second layer appeared at the atop position of the first C60 monolayer. Figure 6b shows a second layer island containing thirteen molecules arranged in a triangular pattern. Similar to the C60 trimer in Figure 6a, all thirteen C60 molecules also occupy the atop position of the first C60 layer, indicating the C60 molecules of second layer adopt the same hexagonal lattice as the first layer. Furthermore, we noticed that the intermolecular spacing for the peripheral molecules is slightly larger than that of the internal molecules. This is different from the second layer of C60 molecules on Au (111) that C60 molecules occupy the three-fold hollow sites of the first layer [46]. It is also different from the simple cubic lattice structure of C60 bulk below 249 K, or the face-centered cubic structure of C60 bulk above 249 K [51]. This phenomenon implies that the charge-transfer induced strong Coulomb interaction among fullerene molecules exist not only in the lateral directions, but also in the normal direction of the Cd substrate.

4. Conclusions

In summary, C60 molecules deposited on the Cd (0001) surface form a variety of domains with different misorientation angles with respect to the lattice directions of Cd (0001). Both orientation disorder and orientation ordering have been observed in the C60 domains. In the (2√3 × 2√3) R 30° domain, orientational disorder of the individual C60 molecules with either pentagon, hexagon, or 6:6 bond facing up has been observed. Orientation ordering appeared in the R26° domain such that all the C60 molecules adopt the same orientation with the 6:6 bond facing up. With the bias variation, possible orientation change of the C60 molecules takes place due to the influence from tip electrical field. In particular, complex chiral motifs composed seven C60 molecules with clockwise or anticlockwise handedness have been observed in the R4° and R8° domains, respectively. Due to the substrate screening and charge transfer from Cd to C60 molecules, the C60 molecules reveal a reduced HOMO–LOMO gap of 2.1 eV, which results in the strong Coulomb interaction among fullerene molecules in lateral and longitude directions. The orientational order and chiral superstructures found in present work provide essential information for understanding the C60-metal interaction, and are of great relevance to the carbon-based nanodevices and nanomaterials.

Author Contributions

Conceptualization, J.W. and J.Y.; methodology, Y.S. and Z.W.; validation, M.T. and K.S.; formal analysis, D.Y. and Y.W.; investigation, Y.S. and Z.W.; resources, J.W.; data curation, M.T. and C.M.; writing—original draft preparation, Y.S. and Z.W.; writing—review and editing, J.W.; supervision, J.W.; project administration, J.W.; funding acquisition, J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China grant number [11874304, 11574253, 11604269, 11674323, 11804282] and the Fundamental Research Funds for the Central Universities grant number (XDJK2020B054).

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant Nos. 11874304, 11574253, 11604269, 11674323, 11804282) and the Fundamental Research Funds for the Central Universities (XDJK2020B054).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stephens, P.W.; Bortel, G.; Faigel, G.; Tegze, M.; Jánossy, A.; Pekker, S.; Oszlanyi, G.; Forró, L. Polymeric fullerene chains in RbC60 and KC60. Nature 1994, 370, 636–639. [Google Scholar] [CrossRef]
  2. Sakurai, T.; Wang, X.D.; Hashizume, T. Scanning tunneling microscopy studies of fullerenes. Prog. Surf. Sci. 1996, 232, 119–154. [Google Scholar]
  3. Yamachika, R.; Grobis, M.; Wachowiak, A.; Crommie, M.F. Controlled atomic doping of a single molecule. Science 2004, 304, 281–284. [Google Scholar] [CrossRef]
  4. Altman, E.I.; Colton, R.J. Determination of the orientation of C60 adsorbed on Au(111) and Ag(111). Phys. Rev. B 1993, 48, 18244–18249. [Google Scholar] [CrossRef]
  5. Lu, X.H.; Grobis, M.; Khoo, K.H.; Louie, S.G.; Crommie, M.F. Spatially mapping the spectral density of a single C60 molecule. Phys. Rev. Lett. 2003, 90, 096802. [Google Scholar] [CrossRef]
  6. Pai, W.W.; Hsu, C.L. Ordering of an incommensurate molecular layer with adsorbate-induced reconstruction: C60/Ag(100). Phys. Rev. B 2003, 68, 121403. [Google Scholar] [CrossRef]
  7. Pai, W.W.; Hsu, C.L.; Lin, K.C.; Sin, L.Y.; Tang, T.B. Characterization and control of molecular ordering on adsorbate-induced reconstructed surfaces. Appl. Surf. Sci. 2005, 241, 194–198. [Google Scholar] [CrossRef]
  8. Li, H.I.; Pussi, K.; Hanna, K.J.; Wang, L.L.; Johnson, D.D.; Cheng, H.P.; Shin, H.; Curtarolo, S.; Moritz, W.; Smerdon, J.A.; et al. Surface Geometry of C60 on Ag(111). Phys. Rev. Lett. 2009, 103, 056101. [Google Scholar] [CrossRef] [Green Version]
  9. Pussi, K.; Li, H.I.; Shin, H.; Loli, L.N.S.; Shukla, A.K.; Ledieu, J.; Fournee, V.; Wang, L.L.; Su, S.Y.; Marino, K.E.; et al. Elucidating the dynamical equilibrium of C60 molecules on Ag(111). Phys. Rev. B 2012, 86, 205426. [Google Scholar] [CrossRef] [Green Version]
  10. Li, H.I.; Abreu, G.J.P.; Shukla, A.K.; Fournée, V.; Ledieu, J.; Serkovic Loli, L.N.; Rauterkus, S.E.; Snyder, M.V.; Su, S.Y.; Marino, K.E.; et al. Ordering and dynamical properties of superbright C60 molecules on Ag(111). Phys. Rev. B 2014, 89, 085428. [Google Scholar] [CrossRef]
  11. Grosse, C.; Gunnarsson, O.; Merino, P.; Kuhnke, K.; Kern, K. Nanoscale Imaging of Charge Carrier and Exciton Trapping at Structural Defects in Organic Semiconductors. Nano Lett. 2016, 16, 2084–2089. [Google Scholar] [CrossRef] [PubMed]
  12. Schull, G.; Berndt, R. Orientationally ordered (7×7) superstructure of C60 on Au(111). Phys. Rev. Lett. 2007, 99, 226105. [Google Scholar] [CrossRef] [PubMed]
  13. Schull, G.; Neel, N.; Becker, M.; Kroeger, J.; Berndt, R. Spatially resolved conductance of oriented C60. New J. Phys. 2008, 10, 065012. [Google Scholar] [CrossRef]
  14. Zhang, X.; Yin, F.; Palmer, R.E.; Guo, Q. The C60/Au(111) interface at room temperature: A scanning tunnelling microscopy study. Surf. Sci. 2008, 602, 885–892. [Google Scholar] [CrossRef]
  15. Gardener, J.A.; Briggs, G.A.D.; Castell, M.R. Scanning tunneling microscopy studies of C60 monolayers on Au(111). Phys. Rev. B 2009, 80, 235434. [Google Scholar] [CrossRef]
  16. Tang, L.; Zhang, X.; Guo, Q.; Wu, Y.; Wang, L.; Cheng, H. Two bonding configurations for individually adsorbed C60 molecules on Au(111). Phys. Rev. B 2010, 82, 125414. [Google Scholar] [CrossRef]
  17. Tang, L.; Zhang, X.; Guo, Q. Organizing C60 molecules on a nanostructured Au(111) surface. Surf. Sci. 2010, 604, 1310–1314. [Google Scholar] [CrossRef]
  18. Tang, L.; Xie, Y.; Guo, Q. Complex orientational ordering of C60 molecules on Au(111). J. Chem. Phys. 2011, 135, 11470211. [Google Scholar] [CrossRef]
  19. Tang, L.; Guo, Q. Orientational ordering of the second layer of C60 molecules on Au(111). Phys. Chem. Chem. Phys. 2012, 14, 3323–3328. [Google Scholar] [CrossRef]
  20. Torrelles, X.; Pedio, M.; Cepek, C.; Felici, R. (2√3 × 2√3)R30 degrees induced self-assembly ordering by C60 on a Au(111) surface: X-ray diffraction structure analysis. Phys. Rev. B 2012, 86, 075461. [Google Scholar] [CrossRef]
  21. Shin, H.; Schwarze, A.; Diehl, R.D.; Pussi, K.; Colombier, A.; Gaudry, E.; Ledieu, J.; McGuirk, G.M.; Loli, L.N.S.; Fournee, V.; et al. Structure and dynamics of C60 molecules on Au(111). Phys. Rev. B 2014, 89, 245428. [Google Scholar] [CrossRef] [Green Version]
  22. Passens, M.; Waser, R.; Karthaeuser, S. Enhanced fullerene-Au(111) coupling in (2√3 × 2√3) R30 degrees superstructures with intermolecular interactions. Beilstein J. Nanotech. 2015, 6, 1421–1431. [Google Scholar] [CrossRef] [Green Version]
  23. Passens, M.; Karthaeuser, S. Interfacial and intermolecular interactions determining the rotational orientation of C60 adsorbed on Au(111). Surf. Sci. 2015, 642, 11–15. [Google Scholar] [CrossRef]
  24. Hashizume, T.; Motai, K.; Wang, X.D.; Shinohara, H.; Saito, Y.; Maruyama, Y.; Ohno, K.; Kawazoe, Y.; Nishina, Y.; Pickering, H.W.; et al. Intramolecular structures of c60 molecules adsorbed on the Cu(111)-(1 × 1) surface. Phys. Rev. Lett. 1993, 71, 2959–2962. [Google Scholar] [CrossRef]
  25. Abel, M.; Dmitriev, A.; Fasel, R.; Lin, N.; Barth, J.V.; Kern, K. Scanning tunneling microscopy and x-ray photoelectron diffraction investigation of C60 films on Cu(100). Phys. Rev. B 2003, 67, 245407. [Google Scholar] [CrossRef] [Green Version]
  26. Schulze, G.; Franke, K.J.; Gagliardi, A.; Romano, G.; Lin, C.S.; Rosa, A.L.; Niehaus, T.A.; Frauenheim, T.; Di Carlo, A.; Pecchia, A.; et al. Resonant electron heating and molecular phonon cooling in single C60 junctions. Phys. Rev. Lett. 2008, 100, 136801. [Google Scholar] [CrossRef]
  27. Wong, S.; Pai, W.W.; Chen, C.; Lin, M. Coverage-dependent adsorption superstructure transition of C60/Cu(001). Phys. Rev. B 2010, 82, 125442. [Google Scholar] [CrossRef] [Green Version]
  28. Xu, G.; Shi, X.; Zhang, R.Q.; Pai, W.W.; Jeng, H.T.; Van Hove, M.A. Detailed low-energy electron diffraction analysis of the (4 × 4) surface structure of C60 on Cu(111): Seven-atom-vacancy reconstruction. Phys. Rev. B 2012, 86, 075419. [Google Scholar] [CrossRef] [Green Version]
  29. Li, G.; Zhou, H.T.; Pan, L.D.; Zhang, Y.; Mao, J.H.; Zou, Q.; Guo, H.M.; Wang, Y.L.; Du, S.X.; Gao, H.J. Self-assembly of C60 monolayer on epitaxially grown, nanostructured graphene on Ru(0001) surface. Appl. Phys. Lett. 2012, 100, 0133041. [Google Scholar] [CrossRef]
  30. Cho, J.; Smerdon, J.; Gao, L.; Suezer, O.; Guest, J.R.; Guisinger, N.P. Structural and Electronic Decoupling of C60 from Epitaxial Graphene on SiC. Nano Lett. 2012, 12, 3018–3024. [Google Scholar] [CrossRef]
  31. Svec, M.; Merino, P.; Dappe, Y.J.; Gonzalez, C.; Abad, E.; Jelinek, P.; Martin-Gago, J.A. van der Waals interactions mediating the cohesion of fullerenes on graphene. Phys. Rev. B 2012, 86, 121407. [Google Scholar] [CrossRef] [Green Version]
  32. Jung, M.; Shin, D.; Sohn, S.; Kwon, S.; Park, N.; Shin, H. Atomically resolved orientational ordering of C60 molecules on epitaxial graphene on Cu(111). Nanoscale 2014, 6, 11835–11840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Monazami, E.; Bignardi, L.; Rudolf, P.; Reinke, P. Strain Lattice Imprinting in Graphene by C60 Intercalation at the Graphene/Cu Interface. Nano Lett. 2015, 15, 7421–7430. [Google Scholar] [CrossRef]
  34. Hou, J.G.; Yang, J.L.; Wang, H.Q.; Li, Q.X.; Zeng, C.G.; Lin, H.; Bing, W.; Chen, D.M.; Zhu, Q.S. Identifying molecular orientation of individual C60 on a Si(111)-(7 × 7) surface. Phys. Rev. Lett. 1999, 83, 3001–3004. [Google Scholar] [CrossRef]
  35. Wang, H.Q.; Zeng, C.G.; Wang, B.; Hou, J.G.; Li, Q.X.; Yang, J.L. Orientational configurations of the C60 molecules in the (2 × 2) superlattice on a solid C60 (111) surface at low temperature. Phys. Rev. B 2001, 63, 085417. [Google Scholar] [CrossRef]
  36. Liu, L.; Liu, S.; Chen, X.; Li, C.; Ling, J.; Liu, X.; Cai, Y.; Wang, L. Switching Molecular Orientation of Individual Fullerene at Room Temperature. Sci. Rep. 2013, 3, 3062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Klyachko, D.V.; Lopez-Castillo, J.M.; Jay-Gerin, J.P.; Chen, D.M. Stress relaxation via the displacement domain formation in films of C60 on Ge(100). Phys. Rev. B 1999, 60, 9026–9036. [Google Scholar] [CrossRef] [Green Version]
  38. Goldoni, A.; Cepek, C.; De Seta, M.; Avila, J.; Asensio, M.C.; Sancrotti, M. Interaction of C60 with Ge(111) in the 3√3 × 3√R30° phase: A (2 × 2) model. Phys. Rev. B 2000, 61, 10411–10416. [Google Scholar] [CrossRef]
  39. Fanetti, M.; Gavioli, L.; Cepek, C.; Sancrotti, M. Orientation of C60 molecules in the, (3√3 × 3√3)R30° and, (√13 × √13)R14° phases of C60/Ge(111) single layers. Phys. Rev. B 2008, 77, 085420. [Google Scholar] [CrossRef]
  40. Leaf, J.; Stannard, A.; Jarvis, S.P.; Moriarty, P.; Dunn, J.L. A combined monte carlo and huckel theory simulation of orientational ordering in C60 assemblies. J. Phys. Chem. C 2016, 120, 8139–8147. [Google Scholar] [CrossRef]
  41. Wang, Y.; Yamachika, R.; Wachowiak, A.; Grobis, M.; Khoo, K.H.; Lee, D.H.; Louie, S.G.; Crommie, M.F. Novel orientational ordering and reentrant metallicity in kx C60 monolayers for 3 ≤ x ≤ 5. Phys. Rev. Lett. 2007, 99, 086402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Stark, R.W.; Falicov, L.M. Band structure and fermi surface of zinc and cadmium. Phys. Rev. Lett. 1967, 19, 795. [Google Scholar] [CrossRef]
  43. Daniuk, S.; Jarlborg, T.; Kontrymsznajd, G.; Majsnerowski, J.; Stachowiak, H. Electronic-structure of Mg, Zn and Cd. J. Phys. Condens. Mat. 1989, 1, 8397–8406. [Google Scholar] [CrossRef]
  44. Edwards, D.A.; Wallace, W.E.; Craig, R.S. Magnesium Cadmium Alloys 4. The cadmium-rich alloys—Some lattice parameters and phase relationships between 25° and 300° structure of the MgCd3 superlattice—Schottky defects and the anomalous entropy. J. Am. Chem. Soc. 1952, 74, 5256–5261. [Google Scholar] [CrossRef]
  45. Tao, M.; Xiao, H.; Sun, K.; Tu, Y.; Yuan, H.; Xiong, Z.; Wang, J.; Xue, Q. Visualizing buried silicon atoms at the Cd-Si(111)-7 x 7 interface with localized electrons. Phys. Rev. B 2017, 96, 125410. [Google Scholar] [CrossRef]
  46. Altman, E.I.; Colton, R.J. Nucleation, growth, and structure of fullerene films on Au(111). Surf. Sci. 1992, 279, 49–67. [Google Scholar] [CrossRef]
  47. Stasevich, T.J.; Tao, C.; Cullen, W.G.; Williams, E.D.; Einstein, T.L. Impurity Decoration for Crystal Shape Control: C60 on Ag(111). Phys. Rev. Lett. 2009, 102, 085501. [Google Scholar] [CrossRef]
  48. Heiney, P.A.; Fischer, J.E.; Mcghie, A.R.; Romanow, W.J.; Denenstein, A.M.; Mccauley, J.P.; Smith, A.B.; Cox, D.E. Orientational ordering transition in solid C60. Phys. Rev. Lett. 1991, 66, 2911–2914. [Google Scholar] [CrossRef]
  49. Kroeger, J.; Neel, N.; Limot, L. Contact to single atoms and molecules with the tip of a scanning tunnelling microscope. J. Phys. Condens. Mat. 2008, 20, 223001. [Google Scholar] [CrossRef]
  50. Lu, X.; Grobis, M.; Khoo, K.H.; Louie, S.G.; Crommie, M.F. Charge transfer and screening in individual C60 molecules on metal substrates: A scanning tunneling spectroscopy and theoretical study. Phys. Rev. B 2004, 70, 115418. [Google Scholar] [CrossRef]
  51. David, W.; Ibberson, R.; Matthewman, J. Crystal structure and bonding of ordered C60. Nature 1991, 353, 147–149. [Google Scholar] [CrossRef]
Figure 1. Initial stage of the self-assembly of C60 molecules on Cd (0001). (a) STM image of an isolated C60 molecule on Cd(0001), 4.5 nm × 4.5 nm, 1.0 V, 28 pA. Inset is the hexagonal lattices of the Cd(0001) thin film grown on the Si(111)-7 × 7 surface, 2 nm × 2 nm, -0.05 V, 35 pA. (b) Step decoration of C60 molecules to a round Cd island, 20 nm × 20 nm, 1.7 V, 25 pA. (c) Close-up view of the C60 chain attaching to the edge of Cd island, 10 nm × 10 nm, 0.5 V, 25 pA. (d) A monolayer island of C60 formed on the Cd substrate, 15.7 nm × 15.7 nm, 2.2 V, 28 pA.
Figure 1. Initial stage of the self-assembly of C60 molecules on Cd (0001). (a) STM image of an isolated C60 molecule on Cd(0001), 4.5 nm × 4.5 nm, 1.0 V, 28 pA. Inset is the hexagonal lattices of the Cd(0001) thin film grown on the Si(111)-7 × 7 surface, 2 nm × 2 nm, -0.05 V, 35 pA. (b) Step decoration of C60 molecules to a round Cd island, 20 nm × 20 nm, 1.7 V, 25 pA. (c) Close-up view of the C60 chain attaching to the edge of Cd island, 10 nm × 10 nm, 0.5 V, 25 pA. (d) A monolayer island of C60 formed on the Cd substrate, 15.7 nm × 15.7 nm, 2.2 V, 28 pA.
Nanomaterials 10 01305 g001
Figure 2. STM images of the (2√3 × 2√3) R30° domain of C60 island on Cd (0001). (a) Uniform domain of the R30° with all C60 molecules in the same contrast except a few dim molecules appeared randomly, 50 nm × 50 nm, 2.2 V, 29 pA. The green and yellow arrows mark the directions of base vector and R30° of the Cd (0001), respectively. (b) Close-up view of the uniform domain without any dim molecules, 10 nm × 10 nm, 1.0 V, 18 pA. The red, green, and yellow circles mark the C60 molecules with a pentagon, hexagon, and 6:6 bond facing up, respectively.
Figure 2. STM images of the (2√3 × 2√3) R30° domain of C60 island on Cd (0001). (a) Uniform domain of the R30° with all C60 molecules in the same contrast except a few dim molecules appeared randomly, 50 nm × 50 nm, 2.2 V, 29 pA. The green and yellow arrows mark the directions of base vector and R30° of the Cd (0001), respectively. (b) Close-up view of the uniform domain without any dim molecules, 10 nm × 10 nm, 1.0 V, 18 pA. The red, green, and yellow circles mark the C60 molecules with a pentagon, hexagon, and 6:6 bond facing up, respectively.
Nanomaterials 10 01305 g002
Figure 3. Bias-dependent STM images (10 nm × 10 nm) of the C60 monolayer in the R26° domain. The bias voltages in (ad) are 1.4, 0.6, −1.4, and −2.0 V, respectively. All the tunneling currents are 30 pA. All the C60 molecules are arranged in the same orientation. The individual C60 molecules reveal sub-molecular contrast such as the two parallel lobes in (a), a small hole and a bright oval in (c), and a bright ball with small opening in (d).
Figure 3. Bias-dependent STM images (10 nm × 10 nm) of the C60 monolayer in the R26° domain. The bias voltages in (ad) are 1.4, 0.6, −1.4, and −2.0 V, respectively. All the tunneling currents are 30 pA. All the C60 molecules are arranged in the same orientation. The individual C60 molecules reveal sub-molecular contrast such as the two parallel lobes in (a), a small hole and a bright oval in (c), and a bright ball with small opening in (d).
Nanomaterials 10 01305 g003
Figure 4. Chiral pinwheel patterns appeared in the R (−4°) and R (+8°) domains. (a) Seven C60 molecules with different directions constitute a chiral pinwheel motif with clockwise handedness in the R (−4°) domain, 10 nm × 10 nm, 2.0 V, 28 pA. There is a central molecule with three-lobe protrusion. (b) Chiral pinwheels with anticlockwise handedness in the R (+8°) domain, 10 nm × 10 nm, 2.2 V, 22 pA.
Figure 4. Chiral pinwheel patterns appeared in the R (−4°) and R (+8°) domains. (a) Seven C60 molecules with different directions constitute a chiral pinwheel motif with clockwise handedness in the R (−4°) domain, 10 nm × 10 nm, 2.0 V, 28 pA. There is a central molecule with three-lobe protrusion. (b) Chiral pinwheels with anticlockwise handedness in the R (+8°) domain, 10 nm × 10 nm, 2.2 V, 22 pA.
Nanomaterials 10 01305 g004
Figure 5. dI/dV spectra acquired on top of a C60 molecule in the (2√3 × 2√3) R30° domain. The HOMO–LUMO gap is 2.12 eV. Tunneling parameters were U = 0.9 V, I = 66 pA before taking the spectra.
Figure 5. dI/dV spectra acquired on top of a C60 molecule in the (2√3 × 2√3) R30° domain. The HOMO–LUMO gap is 2.12 eV. Tunneling parameters were U = 0.9 V, I = 66 pA before taking the spectra.
Nanomaterials 10 01305 g005
Figure 6. C60 clusters on top of the first layer. (a) A C60 trimer appeared in the second layer with each molecule located at the atop site, 2.6 V, 10 nm × 10 nm, 25 pA. (b) A large cluster with thirteen molecules in the second layer, 2.0 V, 6 nm × 6 nm, 30 pA. The individual molecules within clusters are essentially located at the atop position of the first layer.
Figure 6. C60 clusters on top of the first layer. (a) A C60 trimer appeared in the second layer with each molecule located at the atop site, 2.6 V, 10 nm × 10 nm, 25 pA. (b) A large cluster with thirteen molecules in the second layer, 2.0 V, 6 nm × 6 nm, 30 pA. The individual molecules within clusters are essentially located at the atop position of the first layer.
Nanomaterials 10 01305 g006

Share and Cite

MDPI and ACS Style

Shang, Y.; Wang, Z.; Yang, D.; Wang, Y.; Ma, C.; Tao, M.; Sun, K.; Yang, J.; Wang, J. Orientation Ordering and Chiral Superstructures in Fullerene Monolayer on Cd (0001). Nanomaterials 2020, 10, 1305. https://doi.org/10.3390/nano10071305

AMA Style

Shang Y, Wang Z, Yang D, Wang Y, Ma C, Tao M, Sun K, Yang J, Wang J. Orientation Ordering and Chiral Superstructures in Fullerene Monolayer on Cd (0001). Nanomaterials. 2020; 10(7):1305. https://doi.org/10.3390/nano10071305

Chicago/Turabian Style

Shang, Yuzhi, Zilong Wang, Daxiao Yang, Yaru Wang, Chaoke Ma, Minlong Tao, Kai Sun, Jiyong Yang, and Junzhong Wang. 2020. "Orientation Ordering and Chiral Superstructures in Fullerene Monolayer on Cd (0001)" Nanomaterials 10, no. 7: 1305. https://doi.org/10.3390/nano10071305

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