Pavel Kocán*a,
Barbara Pieczyrakb,
Soshiro Umachic,
Martin Cigánekd,
Pavel Sobotíka,
Ivan Ošťádala,
Leszek Jurczyszynb,
Jozef Krajčovič
d and
Kazuyuki Sakamotoc
aFaculty of Mathematics and Physics, Charles University, Prague, Czech Republic. E-mail: pavel.kocan@mff.cuni.cz
bInstitute of Experimental Physics, University of Wroclaw, Wroclaw, Poland
cDepartment of Applied Physics, Osaka University, Osaka, 565-0871, Japan
dBrno University of Technology, Faculty of Chemistry, Materials Research Centre, Purkyňova 118, 612 00, Brno, Czech Republic
First published on 12th February 2025
Electrons confined within the 2D layer of metals grown on silicon substrates exhibit exotic properties due to strong correlation effects. Their properties, such as their 2D superconductivity, have been frequently subjected to possible tuning by doping using charge transfer from adsorbed layers. Doping relies on adding electrons or holes to the system and the resulting shift of the Fermi level EF in the otherwise robust surface electronic structure. This strategy has not been sufficiently controlled in the case of an indium double layer grown on the Si(111) surface. This study provides an alternative approach relying on spatially periodic modification of the surface electronic structure of the 2D metal. Derivatives of diketopyrrolopyrroles (DPP) are used for the growth of perfectly ordered 1D-like molecular superstructures on top of the In double layer, imaged by scanning tunneling microscopy. The integral changes of electronic structure are measured by angle-resolved photoelectron spectroscopy and density functional theory calculations show local modification of the surface states near EF by the adsorbed molecules. This study demonstrates that the surface electronic states can be controllably patterned, using a proper bonding scheme. It is anticipated that the combination of the original 2D superconductor and the 1D-like patterning will motivate further research.
New conceptsWe provide a strategy to confine the electronic states of a 2D metal layer by interaction with self-ordered 1D molecular nanostructures. The molecules create on the surface well ordered molecular wires with a length-to-width ratio of ∼103 driven by interaction via pairs of hydrogen bonds, as observed by scanning tunneling microscopy. Growing structures with desirable effects can be controlled by a choice of organic molecules and deposition conditions. Angle-resolved photoelectron spectroscopy shows that the electronic interaction of the surface with the molecules is gentle-enough to preserve the overall electronic structure. Simultaneously, the density of states of the metal layer at positions of the adsorbed molecules is modified according to density functional theory calculations. So far, the modification of the electronic states of the 2D metals was considered only by net doping – by a charge transfer from the adsorbed layer of the molecules. Here, we show that long-range ordered 1D molecular structures affect the density of 2D metallic states locally. This likely brings new functionalities, as is usual when decreasing the dimensionality of the system. The anisotropic patterning of the 2D electronic states creates a new class of materials, in which a 2D electron gas is interacting with the 1D modulation of the molecular adsorbate. |
Layers of 5p or 6p metals (In, Bi, Tl or Pb) grown under ultra-high vacuum (UHV) conditions on silicon substrates show physical properties of 2D metals. These multivalent elements can both saturate surface dangling bonds with pz orbitals and create 2D metallic states via px and py orbitals. Their rich electronic properties arise from a complex SOC. A 2D electron gas often forms in these materials (,3
,4
5) providing highly mobile electrons with a parabolic dispersion relation in the k-space.
In addition, Rashba-type spin splitting appears in these 2D metals due to the high electric fields present in the near-surface region. This effect was observed by angle-resolved photoemission spectroscopy (ARPES) and explained by density functional theory (DFT) for ,6 Si(111)–Tl(1 × 1)2,7 or Si(110)–Tl(1 × 1).8
For the surface used in this study,9,10 2D superconductivity has been reported11,12 and later explained by dynamic spin-momentum locking, which protects breaking of the Cooper pairs.13 There were attempts to increase the superconducting transition temperature TC ∼ 3 K by hole doping from an extra layer of organic molecules on top of the atomic layer. A nontrivial behavior was observed for phthalocyanine layers on top of the
surface.6,14,15 Depending on the coordinated metal atom of the phthalocyanine, TC can be modified to lower or higher temperatures.6,14 With increasing amount of copper phthalocyanines on the surface, the TC at first increased to decrease again above the full layer,6 accompanied by increasing and decreasing hole doping as observed by ARPES.14 Deposition of the 3,4,9,10-perylenetetracarboxylic dianhydride (PTCDA) molecules with even higher electron affinity surprisingly led to lower TC, which was explained by insufficient ordering of the molecular layer.16
The possibilities of inter-molecular bonding are rather limited on 2D metal layers. Covalent bonding17 via Ullmann coupling or metal coordinated bonding18 typically used on coin-age metals are complicated to realize on 5p or 6p 2D metals, because of the lack of catalytic properties of these metals. The temperature stability of 2D metals may limit thermal activation of the molecular precursors. For example, indium begins to desorb at ∼400 °C from the Si(111) surface, forming different non-2DEG structures. Therefore, we have chosen the stabilization of superstructures by intermolecular hydrogen bonds without the need for thermal activation.19
In this paper, we focus on interaction of the surface with strongly one-dimensional molecular superstructures formed by functionalized diketopyrrolopyrrole (DPP) molecules. We go behind the simplification of the band doping and study the influence of the molecular superstructures on a local level.20–22 We demonstrate that by the local interactions, the pattern of molecular assemblies23 can be inprinted to the 2D metal layer.
DPP-based derivatives are pigments used as organic semiconductors24 in many fields of organic electronics, such as organic field-effect transistors or organic photo-voltaic devices.25 The DPP derivatives are highly versatile due to a wide range of their possible functionalization influencing its opto-electronic properties. The derivatives used here (Fig. 1, full chemical structures are in ESI†) can form 1D chains connected by pairs of hydrogen bonds, as demonstrated at the liquid–solid interface with the molecules in a solution.26 We use scanning tunneling microscopy (STM) to investigate the influence of molecular side groups on their self-ordering and ARPES to quantify changes in the electronic structure of the atomic layer. Details of the interaction are further analyzed by DFT calculations.
![]() | ||
Fig. 1 Derivatives used in the study. Diphenyl- (DP-), dithienyl- (DT-), dithieno[3,2-b] thienyl- (TT-) and dibithienyl- (BT-) diketopyrrolopyrrole (DPP). |
The morphologies of the structures prepared by deposition of DT–DPP molecules are shown in Fig. 2e–h. Even at very low coverage (Fig. 2e), long single-molecule wide chains are observed. The inset in Fig. 2e shows the details of one chain, revealing the corrugation corresponding to single molecules. With increasing coverage, the structure of the chains remain the same, but their density increases. On each atomic terrace, the wires order in parallel to allow the longest possible chains. Fig. 2h shows details of the surface almost fully covered by DT–DPP molecules.
Fig. 2i and j show the ordered molecular structures prepared by deposition of TT– and BT–DPP, respectively. In both cases, the scheme remains the same as in the case of DP–DPP molecules, and only the periodicity reflects the different size of the molecules.
Changes of electronic structure of the layer caused by molecular adsorption were detected by ARPES in the case of representative DP- and DT–DPP molecules. The results are summarized in Fig. 3. Schematic representation of the Brillouin zones (BZs) is depicted in Fig. 3a. The
BZ is marked in green and the 1 × 1 BZ of the idealized In top layer, which is responsible for the superconductivity, is marked by red rectangles. The radius of the circular feature of the Fermi surface (FS) originating from nearly free electrons in the top In layer (marked by violet circles in Fig. 3a) can be used as a measure of hole doping of the layer: with more holes provided to the system, the parabolic band of 2D free electrons shifts up, resulting in a decrease of radius of the circular feature at FS. Fig. 3b–d show the FS measured in the area marked by a blue rectangle in Fig. 3a, for the pristine
surface (b) and for a full adsorbed ML of DP–DPP (c) and DT–DPP (d). The change of the radius upon adsorption ΔkFS can be obtained from fitting momentum distribution curves (MDC) calculated by summation of the photoelectron intensity within the window 0 ± 0.025 Å−1 displayed in Fig. 3e. The colors of MDC in Fig. 3e correspond to the colors of the open circles in Fig. 3b–d indicating the positions of the circular features at FS. By fitting the MDCs using Lorentzian curves, shifts of 0.0022 ± 0.0004 Å−1 and 0.010 ± 0.002 Å−1 were obtained for DP–DPP and DT–DPP molecules, respectively. Fitting the 4d core-level spectra (Fig. 3f) provides another way to obtain the level of doping. In the case of DP– and DT–DPP full monolayers, the energy shifts compared to the pristine surface were −15 ± 7 meV and 23 ± 7 meV, respectively. Compared to the shift of 40 meV in the case of copper phthalocyanine,6 this again shows that the hole doping is insignificant in the case of DP–DPP, but becomes reasonable in the case of adsorption of DT–DPP. The number of transferred carriers can be obtained from the change of radius of the Fermi surface by comparing the size of the Brillouin zone and the change of wave number. The resultant transferred holes per In atom are 0.005 ± 0.001 for DP–DPP and 0.024 ± 0.005 for DT–DPP molecules, respectively.
Structural properties of the molecules adsorbed on the surface and their effect on the local electronic properties of the layer have been investigated using DFT. A relaxed structure of the surface fully covered by DT–DPP molecules is shown as top- and side-views in Fig. 4a and b. The structure is stabilized by a pair of O⋯HN hydrogen bonds. To evaluate this interaction, we have calculated several models with varying number of molecules in the case of DP– and DT–DPP molecules. The resultant hydrogen-bonding energies are −0.70 and −0.71 eV per pair of hydrogen bonds and the bonding energies to the substrate in the case of isolated molecules are −1.81 and −1.56 eV, respectively (for detailed analysis see the ESI†). The interaction energy between DPP molecules, crucial for formation of the linear chains, has been previously calculated using graphite as the substrate. Fu et al.26 obtained a somewhat higher value of ∼ −0.9 eV, probably due to a different competition between an ideal hydrogen bonding configuration and the optimal interaction with the surface.
Cyan and yellow colors in Fig. 4a and b represent electron density depletion and accumulation, respectively, caused by adsorption of the molecules. The charge reorganization is strongly localized to the molecules, with charge depletion (light-blue color) near the topmost In layer and complex reorganization within the molecule. The plane-integrated charge transfer is shown in Fig. 4c. Similar to other systems,14,27 there is a net charge transfer from the area near the metal layer partially towards the molecules and partially towards the metal layer.
As the changes of electronic structure shown in Fig. 4a and b are rather complex, we proceed by plotting the local density of states (LDOS) projected to the top-most In atoms, as affected by adsorbed molecules compared to the pristine layer, see Fig. 4d. To make the projections as comparable as possible, we used large slabs with a single adsorbed molecule of DPP derivatives shown in Fig. 1 per slab. The plotted LDOS are averaged over all In atoms in the topmost In layer. The electronic structure is indeed modified, with all DPP molecules showing a similar effect, somewhat more pronounced in the case of DT–DPP. For additional projections to atoms of molecules, see Fig. S4 in the ESI.† Fig. 4e and f show a comparison of LDOS projected to selected atoms affected and non-affected by the adsorbed DT–DPP molecule. In Fig. 4e and f the selected atom is directly under the central part of the DT–DPP molecule and in the area far from the molecule, respectively. The corresponding In atoms are marked in Fig. 4g by blue and green arrows. In each case, the LDOS is compared to that of the same In atom in the slab without adsorbed molecules (note that the LDOS of In atoms in the clean
shows some variations due to its different positions with respect to the unit cell). Evidently, while the LDOS is preserved on the unoccupied area, it changes significantly under the molecule. The dominant occupied state at ∼ −0.8 eV and the state at the Fermi level shift by ∼0.2 eV towards empty states. The empty state at ∼1.5 eV is significantly reduced. The changes correspond to the complex charge redistribution presented in Fig. 4a and b. Projection of the LDOS in Fig. 4e and f to s- and p-orbitals (Fig. S5 of ESI†) shows that most changes due to adsorption of DPP molecules are related to s- and pz-orbitals. Additional LDOS curves for other studied DPP derivatives are shown in the ESI.†
Next, we discuss how the 1D chains of ordered molecules influence the surface electronic structure. Using DFT calculations, we have demonstrated that quasi-isolated molecules influence the electronic states locally. We propose that this effect would be similar in the case of the 1D chains, such as those in Fig. 2e. DFT calculations of slabs containing rows of molecules together with a bare surface would require extremely large supercells (∼2 × 103 atoms) and are behind recent possibilities. However, we can predict that the influence of individual molecules (see Fig. 4) will combine to create a 1D strip of modified electronic states spatially linked to the chain of molecules. The fine 1D states could be in principle detectable by ARPES under special conditions (single-domain surface, low temperature, high sensitivity), which will be left as a challenge for future studies.
In this study, we have examined four DPP derivatives. Even though these molecules have different energies of HOMO and LUMO states, the changes in the local density of states of the surface are comparable. On the other hand, the side groups influence the molecular assembly on the surface. DT–DPP molecules form isolated 1D chains, and the others create elongated anisotropic islands with periodicity dictated by the size of the molecules. As a result, the different side groups do not change the electronic effect, but rather the spatial distribution of this modified state.
To reproduce the molecular adsorption system considered in the experimental part of our study, the slab approach has been applied. The Si(111) part of the whole substrate was described by the slab composed of six silicon atomic layers. The bulk lattice constant obtained from the minimum energy condition of the bulk unit cell equals 5.469 Å. The dangling bonds associated with the silicon atoms from the bottom atomic layer were saturated by hydrogen atoms. During the relaxation of the atomic position of the system, the silicon atoms from the topmost four atomic layers were allowed to relax while silicon atoms from the two lowest silicon layers were frozen in their bulk-like configuration. The topmost part of the Si(111) system was covered by a double layer In system, which forms a
rectangular structure as reported.40 In atoms from the lowest In atomic layer were located directly above the topmost silicon atoms (on-top position), while In atoms from the higher In layer were situated in equal distance from the nearest underneath In atoms (T4 position). All calculations presented in the present paper have been performed with the use of a 12 × 6 unit cell. Because of the large size of the applied unit cell, the corresponding surface Brillouin zone was sampled only by single Gamma point. The cutoff energy applied in all calculations was equal to 450 eV.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nh00622d |
This journal is © The Royal Society of Chemistry 2025 |