Growth of Germanene on Al(111) Hindered by ... - ACS Publications

None of the techniques used up to now have the capability to detect unambiguously the presence of substrate atoms within the ultra-thin film, i.e., se...
0 downloads 0 Views 6MB Size
Subscriber access provided by UNIV OF LOUISIANA

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Growth of Germanene on Al(111) Hindered by Surface Alloy Formation Emanuel A. Martínez, Javier Daniel Fuhr, Oscar Grizzi, Esteban Alejandro Sanchez, and Esteban D. Cantero J. Phys. Chem. C, Just Accepted Manuscript • Publication Date (Web): 06 May 2019 Downloaded from http://pubs.acs.org on May 6, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Growth of Germanene on Al(111) Hindered by Surface Alloy Formation Emanuel A. Martínez§, Javier D. Fuhr†‡,Oscar Grizzi†‡, Esteban A. Sánchez†‡, and Esteban D. Cantero*†‡ §Instituto

†Centro

Balseiro, Universidad Nacional de Cuyo, Argentina.

Atómico Bariloche, Comisión Nacional de Energía Atómica (CNEA), S. C. de Bariloche, Argentina.

‡Consejo

Nacional de Investigaciones Científicas y Técnicas (CONICET), Argentina.

CorrespondingAuthor *Esteban D. Cantero, Centro Atómico Bariloche, Av. Bustillo 9500, S.C. de Bariloche, Río Negro, Argentina. Ph: +54 294 4445100 5392. Email: [email protected]

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 43

Abstract

Obtaining pure group IV 2D films on well behaved substrates is at present a major goal in material science and of great interest for the associated industries. This goal still represents a challenge in surface science since often these materials tend to form alloys; as a consequence some of the proposed 2D films resulted in topics of controversy regarding the top layer elemental composition and the interpretation of the honeycomb patterns measured by STM. Very recently germanene on Al(111) was proposed as a system having a larger gap than silicene and a quantum-spin Hall effect. This system was studied by several techniques including Scanning Tunnel Microscopy, Low Energy Electron Diffraction, Photoemission and Density Functional Theory. None of the techniques used up to now have the capability to detect unambiguously the presence of substrate atoms within the ultra-thin film, i.e., separated from the corresponding substrate, thus leaving open the question of composition or purity of the layer. Here we follow previous guidelines to grow a Ge film on Al(111) with the expected 3x3 arrangement that was assumed to be characteristic of germanene and then we study “in situ” the properties of the films with ion scattering and recoiling spectrometry; a technique particularly suited for determining the elemental composition of the last surface layer. Our results unambiguously show the formation of a mixture of well-ordered Ge and Al atoms, for all the temperatures and conditions tested, in clear disagreement with the pure single germanene layer proposed in previous works. These conclusions led us to investigate by DFT calculations other possible structures compatible with our present results and the previously reported ones. The most favorable alloyed structures obtained by DFT were then compared with new I-V Low Energy Electron Diffraction curves and from this comparison a top surface model composed of 5 Ge atoms and 3 Al atoms is proposed to replace the germanene model.

ACS Paragon Plus Environment

2

Page 3 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Introduction

Expanding the family of 2D materials1 beyond the well characterized graphene is at present a major goal in material science which demands the combined effort of different synthesis approaches, surface characterizations and calculations. Free standing one monolayer sheets are readily achievable for graphene, for some layered materials (hexagonal boron nitride, dichalcogenides), and for some organic molecules2-4; but no other group IV 2D material has been reported yet with such highly desired property. As an alternative, there is a relatively large family of monolayer films of Si, Ge, Sn, that are grown on different substrates and for which some of the typical properties expected for 2D systems were reported, i.e., a high electron mobility, quantum spin Hall effect, Dirac cones near K points, and even the possibility of manufacturing a transistor5. Although the number of publications proposing novel and promising systems in this field has been increasing strongly during the last five years, the massive application of these systems into real devices is still conjectural, waiting for a more complete understanding and experimental characterization of the basic properties, i.e., the atomic and electronic structure, the interaction with the underlying surface, and the effect of adsorbing other elements. As a sign that this field is not fully developed yet there is an increasing number of controversies regarding the formation of alloyed surfaces, the migration of adsorbates into the bulk or from substrate into the top layer, and also regarding the interpretation of the Scanning Tunneling Microscopy (STM) patterns6-8. Among the new 2D systems germanene has received considerable attention in the last few years because according to Density Functional Theory (DFT) calculations it should present a larger gap than silicene and some specific and useful optical properties. As monolayer germanene does not exists as free standing then all the work done at present is for germanene

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 43

grown on a supporting substrate9-22 or hydrogenated and synthetized from CaGe223, 24. For the first case, the role of the substrate is particularly relevant since often it determines the symmetry of the 2D system on top, the buckling of the layer, the charge transfer between film and substrate, and other properties. Obtaining a full characterization of the 2D layer normally requires a combination of techniques. Low Energy Electron Diffraction (LEED) and STM are probably the most frequently used techniques to study the symmetry and the atomic structure, but they are not sensitive to element type. Other element sensitive techniques like X-ray Photoelectron Spectroscopy (XPS) are quantitative but they lack the surface sensitivity to unambiguously delineate the presence of substrate atoms at the very top from those of the underlying substrate (or substrate atoms that suffered displacements). Photoelectron diffraction and holography have a strong potential since they combine the crystallographic precision of LEED plus composition analysis as has been demonstrated for 2D systems in ref.25. Ion based techniques are less frequently used even though they are sensitive to elements and can have a very high top layer resolution. As a drawback, and depending on the mode of operation, ions can produce surface damage and quantification can be affected by shadowing or focusing effects. Within the ion scattering techniques it has been shown that the combination of a forward scattering geometry to detect direct recoils and time-of-flight techniques (TOF-DRS) results in negligible damage and has a very high sensitivity to top layer atoms26-28. In this work we combine LEED, TOF-DRS, Electron Energy Loss Spectroscopy (EELS) and DFT to study the adsorption of Ge on an Al(111) substrate. We follow the adsorption recipes suggested in previous works16-18 and use the saturation of TOF-DRS signals and the disappearance of the Al surface plasmon to determine the complete formation of the monolayer. LEED analysis and the shape of the Ge 3d XPS peak agree with previous publications16-18.

ACS Paragon Plus Environment

4

Page 5 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Unexpectedly, TOF-DRS used with different projectiles provides clear proofs of the presence of Al atoms within the Ge layer grown on top for a broad range of film thickness, in clear contradiction with the expected formation of a pure germanene layer reported in refs. 16-18. Based on these results we performed DFT calculations using the (3x3) symmetry determined by LEED and a top layer composed of both Al and Ge atoms inferred from TOF-DRS. Several models including also the previously proposed ones as germanene were used in dynamic LEED calculations and contrasted with LEED I-V curves measured for coverages of circa one monolayer. The finally proposed structural model is composed of 5 Ge atoms and 3 Al atoms in the unit cell of the top layer. The dynamic LEED calculations for this model provide better agreement with the experimental I-V curves than previous models based on a pure germanene termination. Moreover, in this model two of the Ge atoms in the cell are shifted outward which is in agreement with previous STM measurements16. The results presented here together with previous ones on Au(111) suggest the necessity of revisiting the systems assumed to be pure terminated Xenes that were traditionally known to form alloys with techniques capable of discerning the presence of substrate atoms in the top layer.

Experimental methods The experiments were performed on an Ultra-High Vacuum (UHV) system working in the low 10-10 mbar range with facilities for Auger Electron Spectroscopy (AES), EELS with monochromatized electrons, Ultraviolet Photoelectron Spectroscopy (UPS), LEED equipped with signal amplification by a single Micro Channel Plate, and TOF-DRS with an observation angle that can be varied continuously from 0o to 60o. For TOF-DRS the UHV chamber is connected to a low-energy (1 to 100 keV) ion accelerator equipped with a switching magnet for

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 43

mass selection and pulsing optics for TOF analysis. The flight paths for emitted recoils and reflected projectiles are 0.76 m for the variable angle detector and 1.76 m for some fixed observation angles. We used Ne+, Ar+ and Kr+ projectiles in the energy range of 4 to 10 keV, with a continuous current of 1 to 5 nA measured by a Faraday cup with ϕ 3 mm aperture. After pulsing the incident fluence is further reduced by almost three orders of magnitude, resulting in negligible damage even after acquiring several successive spectra of the same sample condition. The beam shape can be adjusted by variable aperture slits using typically a size of 1 mm. More details of the instrument can be found in ref.28. For comparison with existent literature some XPS spectra were acquired in a separate setup. The sample consisted of a commercial Al(111) single crystal, which was cleaned and prepared by cycles of 1.5 keV Ar+ sputtering followed by annealing at 400oC. The Ge deposition was performed “in situ” by means of an evaporator consisting of a BN crucible inserted in a Ta cover that is heated by electron bombardment. The evaporation rate was maintained in the order of 3 to 4 monolayers per hour. During evaporation, the pressure raised to a few 10-9 mbar. The sample temperature Tsample during evaporations was kept in the range 100oC to 140oC, where the LEED patterns were sharper. This temperature range lies between those reported in ref. 16 and 18. The sample was mounted on a goniometer with three positional axes, rotation capacity around the azimuthal direction (azim) and rotation of the sample normal (inc) in the plane formed by the incoming beam direction and the detector direction. Grazing incidence corresponds to inc= 0o while incidence normal to the surface corresponds to inc= 90o. The azimuthal angle was labeled as azim=0o when the sample was oriented with the [101] direction contained in plane of the incoming beam and the detector. A schematic representation of the experimental geometry is presented in Fig. 1.

ACS Paragon Plus Environment

6

Page 7 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1. Angle definition in TOF-DRS experiments. The incident beam direction, the detector and the normal to the sample surface are contained in the same plane.

Theory The DFT calculations have been carried out within the slab-supercell approach by using the Vienna “Ab initio” Simulation Package (VASP)29, 30. The one-electron Kohn-Sham orbitals are expanded in a plane-wave basis set and electron-ion interactions are described through the PAW_PBE pseudo-potentials31, 32. Exchange and correlation (XC) are described within the PBE functional33. The sampling of the Brillouin zone is carried out according to the Monkhorst-Pack method34. The chosen cut-off energy is 400 eV, electron smearing is introduced following the Methfessel-Paxton technique35 with σ = 0.1 eV and all the energies are extrapolated to zero absolute temperature. The convergence of the energy is kept always on the order of 10−4 eV. For each configuration we performed a full relaxation, except for the two bottom layers, while the forces are assured to be lower than 10−1 eV/nm.

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 43

For bulk Al, by using a cut-off energy of 400 eV and a 24×24×24 k-point mesh we obtained a lattice parameter aAl of 0.40397 nm. This result is in excellent agreement with the experimental value of 0.40495 nm at room temperature. The adsorption of Ge onto a Al(111) surface was studied considering a slab of six layers of Al(111) in a (3×3) unit cell representing the substrate, on top of which we considered different possible structures of a Ge or Ge-Al layer. The vacuum layer region between consecutive slabs is 1.3 nm, thick enough to ensure negligible interactions between periodic images normal to the surface. The adsorption energy per Ge atom was calculated as:

𝐸ads = ―[

𝐸(𝐺𝑒 ― 𝐴𝑙(111)) ― 𝐸(𝐴𝑔(111)) ― 𝑁𝐴𝑙𝐸(𝐴𝑙 𝑏𝑢𝑙𝑘) 𝑁𝐺𝑒

―𝐸(𝐺𝑒 𝑎𝑡𝑜𝑚)]

(1)

Where E(Ge-Al(111)) and E(Al(111)) are the total energy for the corresponding Ge/Al(111) structure and the clean Al(111) respectively. 𝑁𝐴𝑙 and 𝑁𝐺𝑒 are the number of Al and Ge atoms, respectively, in the top layer. For the adsorption energy calculation we took bulk energy E(Al bulk) as reference for Al atoms, in the case of an alloyed surface, and the isolated Ge atom energy E(Ge atom). The LEED calculations were performed using the AQuaLEED package36, which is based on the Barbieri/Van Hove SATLEED package37. Results and discussion a) Growth of Ge films by evaporation: The growth of Ge was firstly characterized using LEED and TOF-DRS. For all the stages of Ge deposition the obtained LEED patterns were mostly consistent with the accepted (3x3) structure16. At some initial stages we could also observe spots with very low intensity

ACS Paragon Plus Environment

8

Page 9 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

characteristic of a ( 7x 7)R19.1o structure as was described before in ref.18. No other type of reconstruction was observed for sample temperatures between 100 oC and 140 oC during the evaporation. Typical diffraction patterns as a function of increasing evaporation times are presented in Fig. 2. After circa 40 min of evaporation the shape of the spots became less sharp and the background intensity increased. According to the calibration of the growth rate discussed below a monolayer was completed in the range of 15 to 20 min. Additional LEED patterns observed during growth and for different energies are presented in Fig. S 1 of the Supp. Inf., while the comparison of the experimental I-V LEED curves with calculated ones for different structural models is treated further below, at the end of the discussion section.

Figure 2. LEED patterns obtained with electrons around 75 eV for different Ge evaporation times.

The fact that during all evaporation stages the diffraction pattern remains as (3x3) is on one hand beneficial in the sense that it ensures that we are dealing with the phase reported in the literature16-18, however it presents the difficulty of not knowing exactly when a monolayer

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 43

becomes complete. XPS measurements performed in a different setup also indicated a continuous growth of the Ge intensity becoming difficult to establish the point of attaining the monolayer with precision. The comparison between the Ge 3d peak measured in this setup with those reported in references17 and 18 presents a good agreement (accepting the different resolution of a commercial XPS system and the synchrotron based one) (Fig. S 3 of Supp. Inf.). To get information about the formation of a single layer we performed a combined TOF-DRS and EELS study performed at low electron energies as a function of evaporation time. We followed the decreasing Al recoil signal coming from the top layer together with the increasing Ge recoil signal. Due to the high surface sensitivity resulting from shadowing and blocking of the ion trajectories, for a continuous layer or Frank-van der Merwe type growth we expect that the substrate signal should disappear completely38 around one monolayer and that of Ge should attain a constant value. To check this, we used the azimuthal direction azim=19o where the Al recoil signal is particularly high, i.e. along an open direction where focusing onto Al atoms results in a high Al recoil peak26. At the same evaporation times we followed the EELS signal corresponding to the Al surface plasmon which should be more affected by the evaporation of Ge than the bulk plasmon (which was also measured, Supp. Inf.). At the electron energies used here (62 eV) the surface plasmon is observed best at the specular condition. In Fig. 3(a) we present TOF-DRS spectra obtained with Ar+ (4.2 keV) projectiles. The main experimental parameters, using the notation described in the previous section, were inc = 13o, azim=19o (recoils scattered in the [312]direction), scat = 30o, Ldet = 0.75 m, Tsample= 120oC. A low scattering angle of 30o was chosen here because with Ar projectiles the Al recoil peak appears at the left side of the high Ar scattering peak where there is no other contribution and can thus be easily separated. The spectra, from top to bottom, correspond to the scattering from the

ACS Paragon Plus Environment

10

Page 11 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

clean Al(111) surface until 29 minutes of Ge evaporation. The peak in the (5- 5.5) s region corresponds to direct recoils of Al, while the Ge recoils produce a peak located about 8.7 s. The strong signal around 6 s corresponds to scattering of projectiles off Al and off Ge.

Figure 3. (a) TOF-DRS spectra obtained with Ar+ (4.2 keV) along the 19o azimuthal direction [3 12] for 13o incidence during Ge evaporation. (b) Energy loss spectra for 62 eV electrons during Ge evaporation. Top spectra correspond to the clean surface; numbers indicate evaporation time in minutes.

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 43

Along the [312] azimuthal direction, as the evaporation time increases the Al peak is reduced while at the same time the Ge peak grows, reaching a stable condition at about 15 min of evaporation, which we assume to correspond to the completion of a first monolayer. In agreement with this behavior the Ar scattering peak shifts towards lower TOF due to the scattering off (heavier) Ge atoms. Careful observation of the Al peaks shows that its intensity decreases but does not disappears completely. This point is important and will be developed further below. In Fig. 3 (b) we show the corresponding EELS spectra, which were obtained using an electron beam of 62 eV impinging at 45o from the sample normal and with the detector collecting the electrons backscattered to the specular direction in a narrow cone of approximately 2o. For the clean surface, a clear peak appears around 10.7 eV of energy loss, which corresponds to the excitation of a surface plasmon in Al39. During evaporation the signal from the Al surface plasmon decays very swiftly, and the energy loss peak shape evolves towards a broader peak, which has contribution of the excitation of the remaining Al bulk plasmon and of a Ge excitation39. The final form of the loss features seems to have components expected for bulk Ge plus some less intense contribution in the lower energy loss region, close to that calculated for germanene in ref. 40. The fast attenuation of the Al surface plasmon can therefore be considered as an indication that a first monolayer is achieved around 15 min of evaporation. We present in Fig. 4 the area (integral of the peaks after background subtraction) of the recoils and plasmon signals as a function of evaporation time, where the behavior described above can be clearly observed. EELS spectra for the evaporated sequence recorded with the electron detector placed along a non-specular direction to see better the bulk Al plasmon are presented in Fig. S 4 of Supp. Inf.

ACS Paragon Plus Environment

12

Page 13 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 4. Integral of the Al and Ge direct recoil and plasmon peaks from spectra of Fig. 3 during Ge evaporation. The areas of the Al DR peak and the surface plasmon peak were normalized to 1 at zero evaporation time. The inset shows an expansion of the DR intensities for long evaporation times.

(b) Coexistence of Ge and Al atoms at the top layer We mentioned above that after 29 minutes of evaporation, an amount of Al recoil signal is still present, without decaying in time (inset of Fig.4), which means that a fraction of Al atoms from the top layer are detected after completion of the first monolayer and beyond. In TOF-DRS, the

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 43

impinging and departing atom trajectories are governed by a combination of shadowing and blocking effects respectively, with some focusing effects at the edges of shadowing and blocking cones which have radius of the order of 0.1 nm. As a result, the particles that reach the detector must come mainly from the outermost surface layer26. This condition is enhanced by choosing heavy projectiles, such as Ar or Kr at relatively low energy. In this section, based on these effects, we present additional proofs of the coexistence of Al and Ge atoms at the top layer and discuss their ratio. First, and in order to illustrate the high sensitivity of TOF-DRS to the crystallographic structure we show in Fig. 5a spectra taken with Ar+ projectiles along three azimuths of the clean (before evaporation) Al surface. Ar scattering is observed with high intensity at the three azimuths while the Al recoil has a lower intensity and presents a strong dependence on azimuth. Along the azimuth corresponding to the shortest interatomic distance (azim=0o) no Al recoils are produced because of blocking, while they can be observed along azim=30o and azim=19o. In particular, for the latest there is a strong focusing by lateral atoms which generates a very high intensity, as is evidenced in panel b of Fig. 5 upon plotting the integral of the Al recoil peaks (green line) measured as a function of azimuthal angle (here plotted in polar coordinates). In this case the variations are easily interpreted from the known crystallography of the clean surface. After 10 min of Ge evaporation, the spectrum along 0o shows a clear increase in the Al recoil intensity (Fig. 6a) that suggests the existence of Al atoms at the top layer; on the contrary, if the layer were composed of bulk Al termination with Ge on top this increase in the Al signal could not happen. A full azimuthal angle scan, obtained as the areas of Al and Ge recoil peaks in TOF-DRS spectra measured at different azimuths is shown in Fig. 6b for approximately full monolayer coverage. Here we observe that the Al recoil signal is present on all the studied angular range, and in

ACS Paragon Plus Environment

14

Page 15 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

particular in the region between -10o and 10o azimuthal angles. From this figure we conclude that at full coverage: a) there are Al and Ge atoms exposed at the top surface, b) the surface is well ordered, c) Al atoms are located in different positions from those of the clean surface, d) the Al atoms at the top are not coming from defects (they are well ordered) and their amount is comparable to that of Ge. These results are more consistent with formation of an alloyed surface.

Figure 5. (a) TOF-DRS spectra obtained with Ar+ (4.2 keV) along three azimuthal directions for 13o incidence on the clean Al(111) substrate. The inset shows an expanded view around the Al recoil peak. (b) Area of the Al recoil peak versus azimuthal angle (in polar coordinates) superposed to the schematics of the top Al(111) surface.

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 43

Figure 6. (a) TOF-DRS spectra obtained with Ar+ (4.2 keV) along the 0o azimuthal direction at 13o incidence for the clean Al(111) substrate and after 10 min of Ge evaporation. (b) Integral of the recoils peak areas divided by the recoiling cross section plotted vs azimuthal angle.

Very often and depending on the combination of projectile type and geometry chosen for the TOF-DRS spectra, recoil peaks can be very small in comparison with the contribution of the projectile scattering from the substrate because of the higher cross section and the many multiple paths that the latter can undergo. This can be worse when the masses involved are such that the recoil peak appears at the tail of the high TOF side, as it happens for Ge recoils excited by Ar. In this case one can opt for using a heavier projectile, such as Kr where the contribution from Kr-Al scattering will not be present at 45o or higher scattering angles. In this case the spectrum is

ACS Paragon Plus Environment

16

Page 17 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

composed of Al and Ge recoils plus some Kr scattering from Ge. Typical spectra for 5 keV Kr+ ions, for two azimuthal orientations of the sample, are shown in Fig.7. Here the geometry was inc = 20o, azim=0o and 30o (recoils scattered along he [101] and [211] directions), scat = 45o, Ldet = 1.76 m. At this condition the cross sections are similar. Before evaporation, the spectra for both azimuthal directions have a clear different shape due to the crystallographic order of the sample. At 0o the Al contribution appears broader, shifted to lower TOF and is composed mainly from multiple collisions, i.e., no single or true direct recoiling should be present along this direction (inset Fig. 7a). At 30o it appears narrow and at the TOF corresponding to the true direct recoil position. As the evaporation takes place, the Ge recoil peak and the peak corresponding to Kr scattered by Ge are seen clearly (Fig.7b). For long evaporation times (30 min) the spectra still show a strong contribution of Al recoils, and what is more important along 0o it shifts to the position of the true direct recoil, a condition that can only be obtained if the surface is reordered having Al atoms exposed at the top layer, confirming the above discussion resulting from Ar excitation. The peak of Al recoils can be followed beyond 50 min of evaporation time and also at grazing incidence angles (Supp. Inf. – Fig. S 5 and 6) which confirms the assumption of a top surface composed of Al and Ge atoms.

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 43

Figure 7. TOF-DRS spectra obtained with Kr+ (5 keV) along the [101] (azimuthal angle = 0º) and [211](azimuthal angle = 30º) directions at 20º incidence for (a) the clean surface and (b) after 30 min evaporation time. Inset: expanded view around the position of the true Al direct recoil.

b) Ratio of Ge and Al atoms at the top layer We mentioned above that interpreting or identifying the recoiling intensities versus the angles is relatively easy when the crystallography of the surface is known (as in the case of the clean surface), but the reverse and more interesting problem of generating a surface crystallographic model from recoil intensities alone can be extremely complex, particularly if there is more than one type of atoms and probably located at different heights. Instead of this approach we performed calculations based on density functional theory to generate possible models that

ACS Paragon Plus Environment

18

Page 19 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

afterwards were contrasted with all the experimental evidence. This is described in the next section. The models to be considered should be consistent with a (3x3) symmetry (obtained from LEED) and contain both Al and Ge atoms. One issue that requires more analysis is the relative amount of Ge and Al at the top layer. TOF-DRS has enough sensitivity to delineate top layer contributions from subsurface contributions but this property comes from the strong shadowing and focusing effects which make quantitative analysis less precise. Note for example that with Ar projectiles along 19o azimuth the Ge recoil intensity is higher than that of Al recoils while with Kr along 0o and 30o azimuths we observe the opposite. In both cases the ratio is within a factor of two. If we choose a lighter projectile we get less shadowing and focusing but we can lose surface sensitivity. A compromise is obtained by using Ne projectiles at somewhat higher energies, around 10 keV. A scan for different azimuthal directions of the sample using 10 keV Ne+ is presented in Fig. 8for the experimental conditions: inc = 13o, scat = 45o, Ldet = 1.76 m, and 30 min of evaporation. We observe that the dependence of the recoiling peaks on the azimuthal angle is lower than with heavier projectiles (like in Fig. 6b), indicating that the influence of shadowing and focusing effects is less important; however it cannot be ruled-out to obtain a precise Ge/Al ratio. Also, the Al peak now appears at the tail of the very strong Ne scattering peak making the integration of the Al recoil more difficult. The areas of the recoil peaks weighted by recoil cross sections give Ge/Al ratios within the range of 0.6< Ge/Al < 1.5. Similar ratios were obtained for larger incident angles, and for a broad range of Ge evaporation times, ranging from submonolayer to the equivalent of several layers, thus revealing a strong relocation of Al atoms within the topmost Ge layer.

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 43

Figure 8. TOF-DRS spectra obtained with Ne+ (10 keV) at 13o incidence for different azimuthal angles after 30 min of Ge evaporation.

c) Possible Ge/Al crystallographic models inferred from DFT Calculations We explored by means of DFT relaxations possible structures compatible with the experimental data presented above. We restrict the calculations to a (3×3) unit cell compatible with the LEED pattern. We first calculated the previously proposed model13-15 with 8 Ge atoms in the top layer corresponding to a (2×2) germanene. The relaxed configuration is shown in Fig. 9 (structure 1) and has an adsorption energy of 4.32 eV per Ge atom. To explore other configurations with only Ge atoms in the top layer, we performed an “Ab Initio” Molecular

ACS Paragon Plus Environment

20

Page 21 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Dynamics (AIMD) simulation at 80 °C starting from structure 1 and identify possible local minima in the energy landscape. In this way we found the structure 2 shown in Fig. 9 which has a higher adsorption energy of 4.36 eV per Ge atom. This configuration can be seen as large 12 atoms Ge-rings, with a Ge-trimer inside each of them. Starting from these two configurations, structures 1 and 2, we considered alloyed structures replacing some Ge atoms by Al atoms. With this procedure we found the structures 3 and 4 of Fig. 9 which have almost the same adsorption energy of 4.43 eV per Ge atom, i.e. 70 meV higher than the best configuration with only Ge atoms. Structure 3 corresponds to a (2×2) honeycomb structure with Ge and Al atoms intercalated with a 1:1 ratio. On the other hand, the structure 4 corresponds to an equal mixture of Ge and Al atoms in the large rings of structure 2, while maintaining the Ge trimer inside them, resulting in a Ge:Al ratio of 5:3. We tried many other configurations starting from structures 1 and 2 and considering other possible Al replacement of Ge atoms. We also performed several AIMD starting from structures 3 and 4 to identify other possible local minimum configurations. None of the obtained relaxed structures had higher adsorption energies. Structure 3 has one Ge atom shifted outwards, while structure 4 has two higher Ge atoms. This is consistent with reported STM images of this phase16 where there are only two brilliant spots in the (3x3) unit cell.

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 43

Figure 9. Relaxed configuration for the highest adsorption energy structures. Structures 1 and 2 correspond to configurations with pure Ge in the top layer, while structures 3 and 4 correspond to Ge-Al alloy in the top layer. Ge atoms are colored in green, cyan and orange depending on the height. Al atoms are colored in dark grey for the underlying layers and light grey for Al atoms in the top layer.

d) Comparison with I-V LEED curves We measured I-V curves for six spots present in the (3×3) phase: two of them are also present in the clean Al(111) surface ((1 -1) and (-1 1)), while the other four are fractional ones ((1/3 -

ACS Paragon Plus Environment

22

Page 23 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1/3), (2/3 -1/3), (-2/3 2/3) and (2/3 -2/3)) that only appear when the (3×3) pattern develops. The identification of fractional spots in the LEED patterns is provided in Fig. S 1 of Supp. Inf. We also simulated corresponding I-V curves for the four configurations described above. We note that the LEED simulations were performed with the atomic configurations obtained after the DFT relaxations, and that no optimization of the structure was done to fit the experimental I-V curves. The same procedure was previously applied to the spots (-11) and (1-1) of the clean surface to check the consistency of the method. This is shown in Fig. S 2 of Supp. Inf. We show in Fig. 10 the comparison between experimental and simulated curves, only for the fractional spots because the other two spots could have contribution from clean Al(111) areas if the Ge overlayer does not cover all the sample. First, there is a clear disagreement with the experimental I-V curves of the simulated ones for the structures which have only Ge atoms in the top layer (structure 1 and 2). This is particularly evident in the (2/3 -1/3) and (2/3 -2/3) spots, where the simulated curves show prominent peaks at high energy (> 100 eV) not observed in the experimental curves, and also the experimental broad peak below 100 eV in the (2/3 -2/3) spot is not reproduced by neither of these two structures. The simulated I-V curves of both alloyed structures, structures 3 and 4, improves the agreement with the experimental ones. The structure 4 is the one with the best agreement particularly in the (2/3 -2/3) spot where the I-V curve corresponding to structure 3 present a peak above 100 eV not shown experimentally. We can conclude that the simulated I-V curves corresponding to structure 4 shows all the main features present in the experimental curves for all the fractional spots.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 43

Figure 10. Comparison between experimental and simulated LEED I-V curves. For each structure the simulated I-V curves (red lines) corresponding to four fractional spots ((1/3 -1/3), (2/3 -1/3), (-2/3 2/3) and (2/3 -2/3)) are compared to the experimental data (black circles).

ACS Paragon Plus Environment

24

Page 25 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Conclusions We studied the adsorption of Ge on Al(111) under UHV conditions and keeping the sample temperature during the evaporation in the range of 100 oC to 140 oC. LEED patterns obtained for different evaporation times show a dominant (3x3) symmetry consistent with previous reports16,18. The use of ion scattering at forward angles in combination with the measurement of the Al surface plasmon allows accurate determination of one monolayer coverage. Independent TOF-DRS measurements carried on with Ne+, Ar+ and Kr+ projectiles indicate the presence of Al substrate atoms within the topmost Ge layer. The Al atoms are positioned at well-ordered sites and their amount is comparable to that of Ge suggesting the formation of an alloyed surface. This observation was verified from submonolayer to a few layers, and for the entire sample temperatures investigated. In consistency with TOF-DRS results, the highest adsorption energy obtained from DFT calculations occurs for a (3x3) cell composed of three Al atoms and five Ge atoms, where two of the later are located slightly above the rest, in agreement with previous STM observations. The proposed DFT model used “as is” in dynamic LEED I-V curves (i.e. without optimization of atomic positions in the LEED calculation), presents a much better agreement with experimental curves than other models based on a pure germanene layer and are consistent with TOF-DRS main features. From this analysis we conclude that the current germanene model proposed for the Ge/Al(111) system should be replaced by another one based on an alloyed surface, like the model obtained in this work. This and our previous results for Ge on Au(111)27 show the necessity of using techniques with high elemental sensitivity to better characterize the role of the substrate in recent studies of 2D materials.

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 43

Supporting Information. We provide supporting information on 1) additional LEED patterns, 2) XPS Ge 3d peak, 4) additional electron energy loss spectra, and 5) TOF-DRS spectra acquired at grazing incidence. ACKNOWLEDGMENT We acknowledge fruitful discussions with Dr. María Luz Martiarena concerning DFT models and ion scattering trajectories, with Dr. Guillermo Zampieri in relation to XPS analysis, and partial support from Universidad Nacional de Cuyo (06/C517), CONICET (PIP 112 201501 00274 CO) and ANPCyT (PICT-2015-2589). REFERENCES (1) Cahangirov S.; Sahin H.; Le Lay G.; Rubio A. Introduction to the Physics of Silicene and other 2D Materials. Lectures Notes in Physics 2017, 930, Springer ISSN 0075-8450, DOI:10.1007/978-3-319-46572-2. (2) Novoselov K.S.; Geim A. K.; Morozov S.V.; Jiang D.; Zhang Y.; Dubonos S.V.; Grigorieva I.V.; Firsov A.A. Electric Field Effect in Atomically Thin Carbon Films, Science 2004, 306, 666-669,DOI:10.1126/science.1102896. (3) Wang J.; Deng S.; Liu Z.; Liu Z. The Rare Two-Dimensional Materials with Dirac Cones. National Science Review 2015, 2, 22-39, DOI:10.1093/nsr/nwu080. (4) Hamoudi H. Bottom-up Nanoarchitectonics of Two-Dimensional Freestanding Metal Doped Carbon Nanosheet. RSC Adv. 2014, 4, 22035-22041, DOI:10.1039/C4RA02846E.

ACS Paragon Plus Environment

26

Page 27 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(5) Tao, L.; Cinquanta, E.; Chiappe, D.; Grazianetti, C.; Fanciulli, M.; Dubey, M.; Molle, A.; Akinwande, D. Silicene Field-Effect Transistors Operating at Room Temperature. Nature Nanotechnology 2015, 10, 227-231, DOI:10.1038/nnano.2014.325. (6) Svec, M.; Hapala, P.; Ondracek, M.; Merino, P.; Blanco-Rey, M.; Mutombo, P.; Vondracek, M.; Polyak, Y.; Chab, V.; Martin Gago, J. A.; Jelinek P. Silicene Versus TwoDimensional

Ordered

(√19×√19)R23.4º/Pt(111).

Silicide:

Atomic

Phys.

and

Rev.

B

Electronic 2014,

Structure 89,

of

Si-

201412(R),

DOI:10.1103/PhysRevB.89.201412. (7) Wang, W.; Uhrberg, R. I. G. Investigation of the Atomic and Electronic Structures of Highly Ordered Two-Dimensional Germanium on Au(111). Phys. Rev. Mater. 2017, 1, 074002, DOI:10.1103/PhysRevMaterials.1.074002. (8) Peng, W. B.; Xu, T.; Diener, P.; Biadala, L.; Berthe, M.; Pi, X. D.; Borensztein, Y.; Curcella, A.; Bernard, R.; Prevot, G.; Grandidier, B. Resolving the Controversial Existence of Silicene and Germanene Nanosheets Grown on Graphite. ACS Nano 2018, 12, 4754– 4760, DOI: 10.1021/acsnano.8b01467. (9) Acun A.; Zhang L.; Bampoulis P.; Farmanbar M.; van Houselt A.; Rudenko A. N.; Lingenfelder M.; Brocks G.; Poelsema B.; Katsnelson M.

I.; Zandvliet H. J. W.

Germanene: the Germanium Analogue of Graphene. J. Phys.: Condens. Matter 2015, 27, 443002, DOI:10.1088/0953-8984/27/44/443002. (10) Wang Y.; Li J.; Xiong, J.; Pan Y.; Ye M.; Guo Y.; Zhang, H.; Quhe R.; Lu J. Does the Dirac Cone of Germanene Exist on Metal Substrates? Phys. Chem. Chem. Phys. 2016, 18, 19451-19456, DOI:10.1039/C6CP03040H.

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 43

(11) Li L.; Lu S.; Pan J.; Qin Z.; Wang Y.; Wang Y.; Cao G.; Du S.; Gao H. Buckled Germanene

Formation

on

Pt(111).Advanced

Materials

2014,

26,

4820-4824,

DOI:10.1002/adma.201400909. (12) Zhang L.; Bampoulis P.; Rudenko A.N.; Yao Q.; van Houselt A.; Poelsema B.; Katsnelson M. I.; Zandvliet H. J. W. Structural and Electronic Properties of Germanene on MoS2. Phys. Rev. Lett. 2016, 116, 256804, DOI:10.1103/PhysRevLett.116.256804. (13) Dávila M. E.; Le Lay G. Few Layer Epitaxial Germanene: a Novel Two-Dimensional Dirac Material. Sci Rep. 2016, 6, 20714, DOI:10.1038/srep20714. (14) Dávila M. E.; Xian L.; Cahangirov S.; Rubio A.; Le Lay G. Germanene: a Novel TwoDimensional Germanium Allotrope Akin to Graphene and Silicene. New Journal of Physics 2014, 16, 095002, DOI:10.1088/1367-2630/16/9/095002. (15) Matusalem, F.; Koda D. S.; Bechstedt, F; Marques, M.; Teles, L. K. Deposition of Topological Silicene, Germanene and Stanene on Graphene-covered SiC Substrates. Scientific Reports 2017, 7, 15700, DOI:10.1038/s41598-017-15610-3. (16) Derivaz M.; Dentel D.; Stephan R.; Hanf M.; Mehdaoui A.; Sonnet P.; Pirri C. Continuous Germanene Layer on Al(111). Nano Lett. 2015, 15, 2510-2516, DOI:10.1021/acs.nanolett.5b00085. (17) Stephan R.; Hanf M. C.; Derivaz M.; Dentel D.; Asensio M. C.; Avila J.; Mehdaoui A.; Sonnet P.; Pirri C. Germanene on Al(111): Interface Electronic States and Charge Transfer. J. Phys. Chem. C. 2016, 120, 1580-1585, DOI:10.1021/acs.jpcc.5b10307.

ACS Paragon Plus Environment

28

Page 29 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(18) Wang W.; Uhrberg R. I. G. Coexistence of strongly buckled germanene phases on Al(111). Beilstein J. Nanotechnol. 2017, 8, 1946–1951, DOI:10.3762/bjnano.8.195. (19) Fukaya Y.; Matsuda I.; Feng B.; Mochizuki I.; Hyodo T.; Shamoto S. Asymmetric Structure of Germanene on an Al(111) Surface Studied by Total-Reflection High-Energy Positron Diffraction. 2D Mater. 2016, 3, 035019, DOI:10.1088/2053-1583/3/3/035019. (20) Fang J.; Zhao P.; Chen G. Germanene Growth on Al(111): A Case Study of Interface Effect. J. Phys. Chem. C 2018, 122, 18669–18681, DOI:10.1021/acs.jpcc.8b03534. (21) Marjaoui A.; Stephan R.; Hanf M. C.; Diani M.; Sonnet P.Tailoring the Germanene– Substrate Interactions by Means of Hydrogenation. Physical Chemistry Chemical Physics 2016, 18, 15667-15672, DOI:10.1039/C6CP01906D. (22) Zhuang J.; Liu C.; Zhou Z.; Casillas G.; Feng H.; Xu X.; Wang J.; Hao W.; Wang, X.; Dou S.X.; Hu Z. Dirac Signature in Germanene on Semiconducting Substrate. Advanced Science 2018, 5, 1800207, DOI:10.1002/advs.201800207. (23) Bianco E.; Butler S.; Jiang S.; Restrepo O. D.; Windl W.; Goldberger J. E. Stability and Exfoliation of Germanane: a Germanium Graphane Analogue. ACS Nano 2013, 7, 4414– 4421, DOI:10.1021/nn4009406. (24) Hanks A.; Jiang S.; Esser B. D.; Goldberger J. E.; McComb D. W. Electron Diffraction of Germanane.

Microsc.

Microanal.

2017,

23,

1744-1745,

DOI:10.1017/S1431927617009382. (25) Kuznetsov M. V.; Ogorodnikov I. I.; Usachov D. Y.; Laubschat C.; Vyalikh D. V.; Matsui F.; Yashina L. V. Photoelectron Diffraction and Holography Studies of 2D

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Materials

and

Interfaces.

J.

Phys.

Soc.

Page 30 of 43

Jpn.

2018,

87,

061005,

DOI:10.7566/JPSJ.87.061005. (26) Rabalais, J. W. Scattering and Recoiling Spectrometry: An Ion's Eye View of Surface Structure. Science 1990, 26, 521-527, DOI: 10.1126/science.250.4980.521. (27) Cantero E. D.; Solis L. M.; Tong Y. Fuhr J. D.; Martiarena M. L.; Grizzi O.; Sánchez E. A. Growth of Germanium on Au(111): Formation of Germanene or Intermixing of Au and Ge Atoms? Phys. Chem. Chem. Phys. 2017, 19, 18580-18586, DOI:10.1039/C7CP02949G. (28) Salazar Alarcón L.; Jia, J.; Carrera, A.; Esaulov V. A.; Ascolani H.; Gayone J. E.; Sánchez, E. A.; Grizzi, O. Direct Recoil Spectroscopy of Adsorbed Atoms and SelfAssembled

Monolayers

on

Cu(001).

Vacuum

2014,

105,

80



87,

DOI:10.1016/j.vacuum.2014.01.017. (29) Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558 – 561, DOI:10.1016/0022-3093(95)00355-X. (30) Kresse, G.; Hafner, J., Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal– Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B 1994, 49, 14251 – 14269, DOI:10.1103/PhysRevB.49.14251. (31) Blöchl, P. E.; Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953 – 17979, DOI:10.1103/PhysRevB.50.17953. (32) Perdew, J. P.; Wang, Y. Accurate and Simple Analytic Representation of the ElectronGas

Correlation

Energy.

Phys.

Rev.

B

1992,

45,

13244

-13249,

DOI:10.1103/PhysRevB.45.13244.

ACS Paragon Plus Environment

30

Page 31 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(33) Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865-2868, DOI:10.1103/PhysRevLett.77.3865. (34) Monkhorst, H. J.; Pack, D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188 – 5192, DOI:10.1103/PhysRevB.13.5188. (35) Methfessel, M.; Paxton, A. T. High-precision Sampling for Brillouin-Zone Integration in Metals. Phys. Rev. B 1989, 40, 3616 – 3621, DOI:10.1103/PhysRevB.40.3616. (36) Lachnitt, J.: AQuaLEED [software]. https://physics.mff.cuni.cz/kfpp/povrchy/software. (37) Barbieri, A. and Van Hove, M. A.. http://www.icts.hkbu.edu.hk/vanhove/. (38) Diebold U.; Pan J. M.; Madey T. E. Ultrathin Metal Film Growth on TiO2(110): An Overview. Surface Science 1995, 331-333, 845-854, DOI:10.1016/0039-6028(95)00124-7. (39) Thirlwell J. The Characteristic Energy Losses of Slow Electrons Reflected fromAluminium, Germanium, Copper and Gold. Proc. Phys. Soc. 1967, 91, 552 – 564, DOI:10.1088/0370-1328/91/3/305. (40) Rast L.; Tewary V. K. Electron Energy Loss Function of Silicene and Germanene Multilayers on Silver. 2013, arXiv:1311.0838 [cond-mat.mtrl-sci].

ACS Paragon Plus Environment

31

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 43

TOC Figure:

ACS Paragon Plus Environment

32

The Page Journal 33 ofof 43Physical Chemistry 1 2 3 4 5 6 Paragon Plus Environment ACS 7 8

(a)

(b)

clean (d)

10' (e)

30'

40'

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(c)

20' (f)

ACS Paragon Plus Environment

50'

Page 34 of 43

(a )

C o u n ts ( a r b . u n its )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

The Journal of Physical Chemistry

A r - A l/G e

A l

(b )

G e (x 5 ) c le a n

1 1 1 2 2

5

2

6

7

3 ' 6 ' 9 ' 2 ' 5 ' 8 ' 1 ' 4 ' 9 '

E le c tr o n in te n s ity ( a r b . u n its )

Page 35 of 43

ACS Paragon Plus Environment

8 T O F ( µs )

9

0

c le a n 3 ' 6 ' 9 ' 1 2 ' 1 5 ' 1 8 ' 2 1 ' 2 4 ' 2 9 '

1 0 2 0 3 0 E le c tr o n e n e r g y lo s s ( e V )

The Journal of Physical Chemistry

Page 36 of 43

1 .4 0 .1 5

1 .2

G e D R

A re a

A r e a ( a r b . u n its )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

0 .1 0

A l D R

0 .0 5

1 .0

1 5

0 .8

2 0

2 5

E v a p o r a tio n tim e ( m in )

0 .6

0 .2 0 .0

x 5

A l D R G e D R P la s m o n

0 .4

0

ACS Paragon Plus Environment

5

1 0

1 5

2 0

E v a p o r a tio n tim e ( m in )

2 5

3 0

Page 37 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43

The Journal of Physical Chemistry

ACS Paragon Plus Environment

The Journal of Physical Chemistry

A r - A l/G e

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

c le a n

G e

A l

Page 38 of 43

A l ( c le a n ) x 0 .2 A l ( 2 0 m in ) G e ( 2 0 m in )

(b )

A r e a ( a r b . u n its )

C o u n ts ( a r b . u n its )

(a )

5

1 0 m in 6

7 T O F

8

9

1 0

-3 0

-1 5

0

A z im u th a l a n g le ( ACS µs ) Paragon Plus Environment

1 5

3 0 (d e g )

C o u n ts ( a r b . u n its )

Page 39 of 43 1 2 3 4 5 6 7 8 9 10

3 0 m in G e e v a p . ( a ) c l e The a n Journal A l ( 1 1 of1 Physical ) ( b ) Chemistry

A l 3 0 º

3 0 º

A l G e

0 º

K r-G e

A l D R

0 º

1 3

1 4

1 5

1 6

T O F (u s )

3 0 º

1 7

A l D R

0 º

ACS Paragon Plus Environment

1 5 2 0 2 5 3 0 3 5 T O F (u s )

1 5 2 0 2 5 3 0 3 5 T O F (u s )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

C o u n ts ( a r b . u n its )

The Journal of Physical Chemistry

G e θa z

3 2 º 2 3 º 1 4 º 5 º -4 º -1 3 º -2 2 º -3 1 º ACS Paragon Plus Environment -4 0 º

A l 1 2

1 6 2 0 T O F ( µs )

2 4

Page 40 of 43

Page 41 of 43 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40

The Journal of Physical Chemistry

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 42 of 43

Page 43 of The 43Journal of Physical Chemistry 1 2 3 4 5

ACS Paragon Plus Environment