Sie sind auf Seite 1von 12

FEATURE ARTICLE

Scanning Tunneling Characterization www.pss-b.com

Growth of PTCDA Films on Various Substrates Studied


by Scanning Tunneling Microscopy and Spectroscopy
Nicoleta Nicoara, Jose M. Gomez-Rodríguez, and Javier Mendez*

Scanning tunneling microscopy (STM)


The application of scanning tunneling microscopy and spectroscopy (STM/ has been extensively used for the study of
STS) to the study of structural and electronic properties of organic/inorganic ordered organic films grown on different
systems is presented here. The investigation of PTCDA (3,4,9,10, perylene substrates, showing their structure and
morphology. In addition to the topographic
tetracarboxylic dianhydride, C24O6H8) on semiconductor or metal surfaces by
images, scanning tunneling spectroscopy
means of STM/STS shows a wide diversity of results indicating a complex (STS) provides information about the
behavior of the PTCDA molecules. It has been found that the surface surface density of states from the current
reactivity highly influences the PTCDA adsorption and growth properties. The versus voltage dependence.
degree of molecule–surface interaction can lead to randomly chemisorbed This article reviews some representative
examples of PTCDA molecular layers
molecules on highly reactive substrates, to one-dimensional row formation
grown and measured under ultrahigh
on moderately reactive surfaces and well-ordered overlayers in the case of vacuum conditions with STM. Results of
physisorbed PTCDA molecules on weakly reactive surfaces. PTCDA growth on several different sub-
strates, including metals, semiconductors,
and two-dimensional materials, are shown,
emphasizing their similarities and differ-
ences arising from the variability of the molecule–substrate
1. Introduction interaction. We present our investigations concerning the
morphology of PTCDA molecules on different substrates,
PTCDA (3,4,9,10, perylene tetracarboxylic dianhydride, metallic, semiconductors and graphene, together with a view
C24O6H8) are small conjugated molecules of the perylene of the electronic properties of these organic overlayers by means
family, which have been intensively investigated in the past of STS images and conductance plots. Finally we show a brief
years. These molecules are a model system with the property of description of metalorganic structures that can be formed with
forming well-ordered layers on many different substrates. PTCDA and iron.
PTCDA films have a semiconductor behavior, with strong
anisotropy in electronic transport and important optical
properties.[1]
2. PTCDA on Metals
Many investigations have been focused on the morphology and
Dr. N. Nicoara
International Iberian Nanotechnology Laboratory
structure of PTCDA on noble metallic substrates: silver,[2–6]
4715-330 Braga, Portugal copper,[7,8] and gold.[9–14] For the case of PTCDA on Au(111), the
Prof. J. M. G
omez-Rodríguez growth mode and the structure of this system have been
Dept. Física Materia Condensada investigated as a function of growth parameters.[12] A layer-by-
Universidad Aut onoma de Madrid layer growth has been observed for the first several layers,[9] and
28049 Madrid, Spain island formation for multilayer coverage.[13] Typically PTCDA
Prof. J. M. G
omez-Rodríguez molecules on most substrates form ordered domains with a
Instituto Nicolás Cabrera herringbone structure, where the unit cell consists in two
Universidad Aut onoma de Madrid
28049 Madrid, Spain molecules in a perpendicular arrangement.[11,12,14] For the
Prof. J. M. G
omez-Rodríguez
Au(111) substrate, also a square structure can be observed.[13]
Condensed Matter Physics Center (IFIMAC) Figure 1a indicates the chemical structure of the PTCDA
Universidad Aut
onoma de Madrid molecule. Figure 1b shows PTCDA molecules covering the
28049 Madrid, Spain Au(111) surface. The two different structures, the less dense
Dr. J. Mendez square structure one with molecules in perpendicular positions
Instituto de Ciencia de Materiales de Madrid, ICMM-CSIC and the rectangular one also named “herring-bone structure” are
28049 Madrid, Spain
E-mail: jmendez@icmm.csic.es
observed on the left and right sides respectively.
On metals, the PTCDA overlayers are characterized by a weak
The ORCID identification number(s) for the author(s) of this article interaction, covering surface steps and surface reconstruction
can be found under https://doi.org/10.1002/pssb.201800333.
without appreciable alteration, as is the case of the Au(111)-
DOI: 10.1002/pssb.201800333 (22  √3) reconstruction. In Figure 2a an STM image measured

Phys. Status Solidi B 2019, 256, 1800333 1800333 (1 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

at RT shows a gold substrate partially covered with PTCDA Nicoleta Nicoara received the
molecules, with the gold reconstruction visible at both sides. Diploma degree in Physics from the
This weak interaction can be also deduced from the spectro- Babes-Bolyai University, Cluj-Napoca
scopic results, hence, STS and UPS show how the gold Shockley (Romania) in 2000 and the PhD
state is barely altered by the PTCDA coverage.[15] The STS degree in Condensed Matter Physics
conductance plots presented in Figure 2b show the Shockley from Universidad Aut onoma de
surface state on clean Au(111) at 0.4 eV, and indicates that for Madrid (Spain) in 2007. During her
1 ML PTCDA coverage this surface state is still visible and that its PhD and postdoc until 2011, she
position in energy is slightly shifted upwards. Such small worked on the nanoscale
alteration of the surface state is only comparable to noble gases characterization of semiconductor
physisorption.[16] materials and also on the development of advanced
At low temperature and using low bias voltages we have been scanning probe microscopy techniques. Since 2012, when
able to visualize electronic standing waves at the PTCDA/ she joined the International Iberian Nanotechnology
Au(111) interface. These electronic patterns are originated from Laboratory (Portugal), her research has been dedicated to
the scattering of the interface state electrons with surface the SPM characterization of the optoelectronic properties
defects.[17] The 2D Fourier transform (FT) of these patterns for of chalcopyrite Cu(In,Ga)Se2 semiconductor materials
energies very close to the Fermi level allowed us to image used in high efficiency thin film solar cells.
simultaneously the Fermi contour of Au(111) and PTCDA/
Au(111) surfaces. These FT show that the Fermi wavevector of
Jose María G omez-Rodríguez is a
the molecular layer is smaller than that of the pristine metal
Professor at Universidad Aut onoma
surface. This can be observed in Figure 3, where STM images
de Madrid (UAM). In 1992, he
acquired at low bias voltages (4 mV) and at 40 K over gold
obtained a PhD in Condensed
partially covered with PTCDA (Figure 3b) show the standing
Matter Physics at UAM, related to
waves. These are visible over the gold areas and due to the weak
scanning tunneling microscopy
PTCDA–gold interaction also over the PTCDA covered areas.
(STM) studies. Then he moved as a
Figure 3c and d show the FT obtained over clean gold areas and
postdoc to the Centre National de la
PTCDA covered ones respectively. Both images show the contour
Recherche Scientifique in Grenoble
of Fermi surfaces corresponding to the Shockley surface state,
(France), where he gained an in-
and indicate that the Fermi wavevector of the molecular layer is
depth knowledge on the study of semiconductor surfaces
smaller than that of the pristine metal surface. As shown by
by STM. Since 2000, he leads the NanoSPM research
Nicoara et al.,[17] from the average radii of the circles shown in
group at UAM in the fields of Nanoscience,
Figure 3(c) and 3(d) the Fermi wave vectors can be determined to
Nanotechnology and Scanning Probe Microscopies
be kF(Au) ¼ 0.17  0.01 Å 1 and kF(PTCDA/Au) ¼ 0.15  0.01
Å 1, for the bare Au(111) surface and for the PTCDA covered
(SPM). Moreover, he was co-founder of the spin-off
company Nanotec Electr onica S.L. (devoted to the design,
one, respectively. As shown in ref. [17], this is in agreement with
development and commercialization of SPMs), and co-
an upward shift in energy of the surface state of the Shockley-
author of the WSxM software, one of the most widely
derived interface state of PTCDA/Au(111) observed both in STS
used programs for SPM data acquisition and processing.
and ARUPS experiments.[15,18]

Javier Mendez studied Physics in the


Universidad Aut onoma de Madrid,
3. PTCDA on Clean Semiconductors Spain, where in 1996 he got his PhD
PTCDA has been considered for years a model system for the degree in Physics honored with
study of its epitaxial growth on metals, as we have seen in the Extraordinary Doctorate Award. He
previous section. In these cases, ordered assemblies are usually performed two postdoctoral stays, at
mediated by hydrogen bonds between the adsorbed molecules, Humboldt University and in the
yielding stable configurations. In semiconductor surfaces, Fritz Haber Institute of the Max
however, due to the existence of dangling bonds, diffusion of Planck Society, both in Berlin,
single molecules is highly hindered and the molecule-substrate Germany. In the second one he
interaction plays a major role on final configurations. investigated catalytic processes in the group of Professor
PTCDA growth has been investigated on pristine silicon and Gerhard Ertl (Nobel prize in 2007). Back in Spain, he
germanium surfaces, like Si(111),[19,20] Si(001)[21,22], Ge(111),[23] joined the Instituto de Ciencia de Materiales de Madrid
and Ge(001),[24,25] as well as on III–V semiconductor compound (ICMM-CSIC) in 2001, obtaining a permanent position in
surfaces, like GaAs(001)[26] and InSb(001),[27,28] among others. 2006. His research subjects are related to surface science
In the present paper, two prototypical examples that have been and nanotechnology, involving organic molecules and
studied by STM are reviewed. The first one deals with PTCDA graphene, characterized with scanning probe
adsorption on Si(111)-(7  7) surfaces.[19] The second one is microscopies (STM, AFM).
related to the adsorption and growth of the first molecular layers
on Ge(111)-c(2  8).[23]

Phys. Status Solidi B 2019, 256, 1800333 1800333 (2 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

Figure 1. a) Chemical structure of PTCDA (3, 4, 9, 10-perylene tetracarboxylic dianhydride). b) STM image (9.6  4.1 nm2) of PTCDA molecules covering
a Au(111) substrate. Two different structures are formed, the square one (left) and the herring-bone PTCDA structure (right). I ¼ 1.3 nA; V ¼ 1.5 V.

The initial stages of PTCDA adsorption on Si(111)-(7  7) non trivial contributions to the STM tunnel current, making it
surfaces was investigated by Nicoara et al.[19] combining essential full first-principles calculations and simulations for the
experimental STM measurements and first-principles DFT correct interpretation of the experimental images.
calculations. Figure 4 shows representative STM images of A second representative example of the situation observed
single PTCDA molecules adsorbed at RT on a Si(111)-(7  7) when adsorbing PTCDA on pristine semiconductor surfaces is
surface. Single molecules are randomly adsorbed on the surface. the case of PTCDA/Ge(111)-c(2  8). In this system, Martínez-
Among them, some present a similar aspect and are adsorbed on Galera et al.[23] have shown by means of STM measurements that
the corner hole of the Si(111)-(7  7) reconstruction. On those on well reconstructed regions of the Ge(111) surface, PTCDA
molecules, bias-dependent imaging allowed to resolve intramo- molecules have a relatively high mobility at RT. This was inferred
lecular features which could not be simply ascribed to the from the experimental observations that showed molecular
frontier molecular orbitals (MO) of the free molecule. Instead, nucleation only in domain walls, steps and surface defects.
quite involved DFT calculations, including realistic STM tips, However, it was not detected any ordering of the molecules for
were needed to understand the shifts and splittings of the MO. coverage below one monolayer. For higher coverage, the
Figure 5 shows the comparison of STM measurements and formation of three-dimensional islands, presenting a typical
first-principles calculations[19] performed on the whole system, herringbone reconstruction, was observed.
Si(111)-(7  7)–PTCDA molecule–STM tip, using a very efficient Figure 6 illustrates the STM observations of PTCDA growth
and accurate method to simulate STM images and spectra on Ge(111)-c(2  8) at submonolayer coverage. For the lowest
developed by Paz et al.[29] As observed in this figure, there is an coverage two dimensional islands separated by Ge(111) regions
excellent agreement between experimental and DFT simulated are imaged. Although there is no ordering within these islands,
images for different bias voltages, both for unoccupied and the bright protrusions are adscribed to PTCDA molecules. This
occupied states. The main conclusions of this work were that the revealed that the molecule–molecule interaction was not enough
strong molecule–substrate interaction in reactive semiconductor to produce a self-assembled molecular layer. Isolated single
surfaces like Si(111)-(7  7) produces broadenings, shifts and PTCDA molecules were, however, not observed on Ge(111). This
splittings of the original PTCDA MO. This fact translates into was interpreted as a sign of a relatively high mobility at RT.

Figure 2. a) STM image of sub-monolayer PTCDA coverage on Au(111). The left side corresponds to the clean substrate. The right side corresponds to a
gold terrace covered with 1 ML of PTCDA. The substrate reconstruction is still visible through the organic layer (57  40 nm2, V ¼ 2.0 V, I ¼ 0.16 nA). b)
Tunneling spectroscopy performed at RT on clean Au(111) and on PTCDA covered gold. For clean gold (dashed line) the onset of the Shockley surface
state is identified around 0.4 eV. A similar peak and the molecular related features HOMO and LUMO appear in the differential conductance spectra
recorded on 1ML PTCDA (black line). Reproduced with permission.[15] Copyright 2006, Elsevier.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (3 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

Figure 3. a) STM image measured in constant current mode at 40 K of PTCDA partially covering the Au(111) substrate (70  70 nm2, V ¼ 1.8 V,
I ¼ 0.02 nA). b) STM image measured at low bias voltage and 40 K showing standing waves on PTCDA and gold areas (V ¼ þ0.004 V, I ¼ 0.02 nA). c) and
d) Corresponding Fourier transforms of the areas marked in image (b). Reproduced with permission.[17] Copyright 2016, IOP Publishing.

Figure 4. a) STM image (11  14 nm2, V ¼ þ1.5 V, and I ¼ 0.27 nA) of the Si(111)-(7  7) surface after exposure to PTCDA at RT. The arrows point to
single molecules adsorbed on the corner hole site of the (7  7)-reconstruction. b) and c) STM images measured on the same 4.2  4.6 nm2 area at
different tunneling conditions: b) V ¼ þ0.4 V, I ¼ 0.3 nA and c) V ¼ 1.5 V, I ¼ 0.26 nA. d) Schematics of the adsorption geometry of PTCDA on Si(111)-
(7  7). Reproduced with permission.[19] Copyright 2010, American Physical Society.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (4 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

Figure 5. This figure compares experimental STM images (top) of PTCDA/Si(111)-(7  7) with first-principles DFT simulations (bottom) for several bias
voltages. In the simulations, STM tips composed of Si atoms were used. Images size: 3.2  3.2 nm2; tunnel current: I ¼ 0.13 nA. Reproduced with
permission.[19] Copyright 2010, American Physical Society.

Figure 6. Large scale (top) and detailed (bottom) STM views of the PTCDA/Ge(111)-c(2  8) system at increasing molecular coverage. Image sizes are
50  50 nm2 for (a), (c), and (e), and 12  12 nm2 for (b), (d), and (f). The sample biases (V) and tunnel currents (I) are: þ1.6 V, 0.1 nA (a); þ1.5 V, 0.1 nA
(b); þ1.5 V, 0.08 nA (c); þ1.5 V, 0.05 nA (d); þ2.0 V, 0.2 nA (e); þ2.0 V, 0.1 nA (f). Reproduced with permission.[23] Copyright, IOP Publishing.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (5 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

Moreover, by using standard nucleation theory arguments, the substrates, surface passivation with Se or with S has been
surface diffusion energy barrier for single PTCDA molecules used.[26,30,31] For silicon, either in situ passivation using atomic
was estimated to be Ed  0.4 eV, consistent with the diffusion of hydrogen or wet chemical etching resulted in atomically flat
single molecules at RT, in contrast with the situation on Si(111)– surfaces on which improved molecular order was
(7  7) where isolated single molecules were indeed observed. obtained.[21,32–34] The following section describes few examples
This situation could be closer to the one found on some metal- where scanning probe microscopy techniques were used to
saturated Si(111) surfaces as Sn/Si(111). obtained insight into structural and physicochemical properties
At coverage between 0.6 and 1 ML (Figure 6c–f), larger of PTCDA layers on semiconductor surfaces, for which
PTCDA islands grow by incorporating new molecules until alternative passivation schemes were used. A comprehensive
coalescence was observed. At coverage above 1 ML (here not overview of PTCDA growth on Pb–Si(111), Sn–Si(111), and S–
shown), the disordered layer act as a passivating layer where 3D GaAs(001) system shows that substrate properties as lattice
PTCDA crystallites can develop,[23] displaying a Stranski– parameter, surface corrugation and reactivity can play an
Krastanov growth type, similar to the situation observed after important role in the formation of ordered molecular layers.
passivation by other methods, as will be shown below.

4.1. PTCDA on Pb/Si(111)


4. PTCDA on Passivated Semiconductors
The inclusion of organic materials in semiconductor devices The growth of PTCDA molecular layers on Pb/Si(111) surfaces
was highly exploited in the recent years due to their has been studied using scanning tunneling microscopy under
optoelectronic properties. However, high performance devices ultra-high vacuum conditions, as reported by Nicoara et al.
require a controlled deposition of the organic films. This is [[35]] The deposition of 1 ML Pb on Si(111) leads to the
sometimes difficult to achieve because semiconductor materi- formation of several different phases with a varying surface
als present a high density of dangling bonds which make their reactivity.[36–38] It has been shown that PTCDA organizes in
surfaces reactive. As shown previously, highly reactive surfaces ordered structures on metallic Pb/Si(111). However, for phases
impede the formation of ordered molecular layers. Therefore with a lower amount of Pb and hence with a higher density of
passivation treatments are usually employed to reduce the active dangling bonds, the molecules adsorb randomly and form
surface reactivity and to improve the ordered growth. For GaAs disordered structures.

Figure 7. STM results of PTCDA on Pb/Si(111) phases. The left panel shows the relation between the Pb coverage, the formed phases and their surface
reactivity. a) STM image of the mosaic phase corresponding to 1/6 ML Pb on Si(111) before and (b) after the deposition of a low PTCDA coverage, which
results in a disordered layer. c) Small size ordered PTCDA domains form on β-(√3  √3)-Pb phase as indicated by the white outlined rectangles. d)
Detail STM image indicating the unit cell of the formed herringbone-like structure. e) Larger PTCDA ordered domains form on (1  1)-Pb phase. f) and g)
Proposed models for a commensurate and a coincident (3  1) PTCDA structure. h) and i) PTCDA ordered layers on HIC phase and the proposed
coincident (3  3) structure model. Reproduced with permission.[35] Copyright 2016, IOP Publishing.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (6 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

Figure 7 shows the evolution of PTCDA growth on Pb/Si the HIC phase, as concluded from the lattice parameters
substrates for which the surface reactivity was systematically and the STM image contrast. The proposed models are shown
modified through the variation of the Pb amount. The different in Figure 7g and i. In addition, the authors proposed a distinct
Pb/Si phases are schematically indicated on the left side of the commensurate structure on the (1  1)-Pb phase. At
figure, whereas representative STM topographic images and the higher coverage (approx. 7 ML) PTCDA grow in a Stranski–
proposed structural models of PTCDA growth on these phases Krastanov mode on (1  1)-Pb and β-(√3  √3)-Pb phases.
are shown on the right side of the figure. It was expected that, by The unit cell of PTCDA within the three dimensional islands
passivating these surfaces, the adsorbate–substrate interaction is comparable to that reported for the β-polymorph of the bulk
would be considerably lower than the case of non-passivated crystal.
silicon surfaces. The variation of the substrate reconstruction in Pb/Si(111)
However, it has been found that the mosaic phase which has system was explored to find alternative passivation methods for
the lowest amount of Pb is still reactive. PTCDA molecules the growth of organic materials on Si(111) surfaces. The study
adsorb randomly and form a disordered layer as shown in allowed finding a correlation between the surface reactivity and
Figure 7b, indicating that the Pb coverage hereof is not sufficient the degree of the structural ordering of PTCDA on these
to passivate the silicon surface. surfaces. STM measurements have shown that low Pb coverage
For Pb coverage between 1/3 and 1 ML, two different Pb/ substrates, as β-(√3  √3)R30 and the mosaic γ-(√3  √3)
Si(111) phases appear at room temperature: the β-(√3  √3)-Pb R30 reconstructions, do not favor the formation of large ordered
and the (1  1)-Pb phases. For the β-(√3  √3)-Pb phase, short PTCDA areas. This was observed instead on the higher Pb
range ordered domains with 10–15 nm2 size were mainly coverage phases, i.e., the hexagonal incommensurate and the
observed, as indicated by the white outlined rectangles in (1  1)-Pb reconstructions. Their inert surfaces permit the
Figure 7c. A higher magnification image, in Figure 7d shows that molecular diffusion at room temperature, thus long range
the molecules develop a herringbone-like structure. The absence PTCDA domains with highly ordered molecules can be obtained.
of larger ordered PTCDA domains was attributed to the high The structural ordering degree of PTCDA on different Pb/
number of defects, which are commonly found on these Si(111) coverage phases was connected with the specific
substrates. It was suggested that a reduction in the number of configuration of Pb atoms on Si(111) surface, which saturate
defects, through changes in the sample preparation, could lead the surface dangling bonds and led to a reduced molecule-
to an increased size of ordered domains. substrate interaction.
Ordered PTCDA layers are observed for high Pb coverage
phases, i.e., the (1  1)-Pb and the hexagonal incommensurate
(HIC) phases, as shown in Figure 7e and h. The authors 4.2. PTCDA on Sn/Si(111)
conclude from the STM results that the diffusion of molecules
on these surfaces favors the formation of packed molecular Yet another approach for semiconductor passivation was
structures even at low PTCDA coverage (0.9 ML). The investigated by adsorbing Sn atoms on Si(111) surfaces,[42]
molecules arrange in a herringbone structure, which is similar to the case of Pb/Si(111) system. It was anticipated that
the prevailing one in PTCDA growth on (111) faces of noble the molecule–substrate interaction would be moderate, i.e., not
metals,[39] or graphene.[40,41] The results suggest that there is a as strong as in the case of clean semiconductors and not as
weak interaction between the PTCDA and the surfaces of weak as in the case of hydrogen passivated surfaces. This was
these phases. The authors proposed two coincident structures previously demonstrated for the Si(111)-(√3  √3)R30 Ag
of the PTCDA overlayer, a (3  1) superstructure formed on surface for which ordered molecular layers were obtained for
the (1  1)-Pb and a (3  3) PTCDA superstructure formed on C60,[43] pentacene,[44] or perylene[45,46] molecules.

Figure 8. STM images of the Si(111)-(2√3  2√3)R30 –Sn surface for successive PTCDA depositions. a) At 0.2 ML, single molecules or chains with few
molecules are observed, V ¼ þ2.0 V, I ¼ 1.0 nA. b) At 0.35 ML, longer and more abundant molecular chains form V ¼ þ2.2 V, I ¼ 0.1 nA. c) At 0.5 ML, the
surface is fully covered by molecular rows, V ¼ þ2.2 V, I ¼ 0.1 nA. The black arrows indicate the substrate direction. Images size: 50  40 nm2.
Reproduced with permission.[42] Copyright 2009, American Chemical Society.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (7 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

Figure 9. STM images of 0.5 ML PTCDA on Si(111)-(2√3  2√3)R30 –Sn surface, acquired at different bias voltage. Intramolecular features (indicated
by blue circles) are comparable to the free molecular orbitals of PTCDA (LUMO and LUMO þ 2 shown in the insets) a) V ¼  2.0 V, b) V ¼ þ1.5 V.
Images size: 8.0  8.0 nm2, I ¼ 1.0 nA. c) Structural model for the PTCDA adsorption geometry: a1 and a2 indicate the substrate lattice vectors. The
dashed line represents the PTCDA long axis. d) Proposed commensurate (4√3  2√3) PTCDA structure on 2√3 Sn surface. b1 and b2 represent the
PTCDA unit cell vectors. Reproduced with permission.[42] Copyright 2009, American Chemical Society.

Figure 10. STM and STS on PTCDA multilayers on a sulfur passivated GaAs(001) substrate. a) STM topographic image of a PTCDA crystal for 14 ML
PTCDA at 180  C (I ¼ 0.1 nA, V ¼ 3.3 V, size 44  44 nm2) and b) STM image showing the unit cell of the PTCDA herringbone structure (10  5.8 nm2).
c) Spectroscopic data on PTCDA. Initial feedback parameters are I ¼ 0.08 nA, V ¼ 3.2 V. Current–voltage plots at different Z distances. Inset: Current–
distance variation indicating an exponential behavior; d) logarithmic representation of the I–V and e) normalized derivative of the I–V in (c). Reproduced
with permission.[26] Copyright 2003, IOP Publishing.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (8 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

The RT deposition of Sn on reconstructed Si(111) leads to the crystals with a herringbone structure, characteristic to PTCDA-
formation of several Sn/Si(111) phases, as a function of the bulk, are resolved in high resolution STM images as shown in
annealing temperature and the amount of Sn on the sample. In a Figure 10a and b.
report by Nicoara et al.,[42] it has been shown by means of STM Scanning tunneling spectroscopy was also performed to
measurements that even the Si(111)-(2√3  2√3)R30 –Sn characterize the electronic properties of the PTCDA molecular
surface maintains its semiconductor character after the Sn layers. The authors have noted that the current–voltage (I–V)
deposition, other aspects such as surface reactivity or substrate characteristics, from which band gap values are estimated, can
periodicity and corrugation can influence the growth of PTCDA. be highly dependent on the tunneling conditions.[47] If the tip-
Figure 8 shows STM topographic images of Si(111)- sample distance is very small, a deformation of the organic layer
(2√3  2√3)R30 –Sn after the deposition of different PTCDA can occur, resulting in distorted band gap values. Therefore the
coverage. The bright protrusions with comparable size and authors proposed monitoring the tunneling current behavior
shape are attributed to individual PTCDA molecules. Already for (i.e., by acquiring current versus distance curves) to establish
the lowest PTCDA coverage the molecules assemble into a chain- optimum conditions for the I–V spectroscopy. With this
like structure. This tendency is confirmed for additional approach they obtained a band gap value of 2.2 eV for PTCDA
exposures. At 0.5 ML (Figure 8c), the molecular chains cover multilayers. They suggested that considerable interaction
uniformly the whole surface, forming an ordered structure with between the molecular films and the STM tip can induce
regularly spaced rows. artifacts in the STS results, which led to inconsistent or
The fact that PTCDA molecules do not assemble in a anomalously low gap values.
herringbone structure, as on most of the inert surfaces,[39] but on
the contrary form a chain like structure was indicative of a
considerable molecule-substrate interaction. By means of high
resolution STM imaging of the PTCDA molecular orbitals 5. PTCDA on Other Substrates: PTCDA on
LUMO and LUMO þ 2 (see Figure 9) the authors are able to Graphene
identify the registry of the molecular structure with respect to the In very recent years, as it could not be otherwise, PTCDA has
substrate lattice (model shown in Figure 9c). It is proposed that been grown also on some of the emerging 2D materials. Most of
the formation of a Sn–O bond could be the driving mechanism to the work has been devoted to its growth on graphene
the formation of this chain-like structure and consequently substrates,[40,41,48] though some other 2D materials as WS2[49]
responsible to the formation of the commensurate (4√3  2√3) and h-BN[50] are starting to be explored. Due to its inert surface,
PTCDA structure at 0.5 ML coverage (model shown in graphene acts as a perfect substrate for the study of self-assembly
Figure 9d). This is one of few examples in which passivation of PTCDA molecules which remain practically undistorted from
processes induce the creation of one-dimensional structures. an electronic point of view with respect to the electronic structure
The system could serve as an organic template for further of the free molecules. An example of such behavior is shown in
functionalization of the semiconductor surfaces. Figure 11, extracted from ref. [41] In that work, Martínez-Galera
et al., studied the molecular and electronic structure of the first
monolayer of PTCDA on graphene grown on Pt(111) surfaces.
4.3. PTCDA on S-GaAs(001) According to Martínez-Galera et al.,[41] PTCDA molecules
present a herringbone structure, analogous to that observed on
Scanning tunneling microscopy and spectroscopy has been used Au(111) surfaces (see above); in this structure the molecules lie
to investigate the structural and electronic properties of PTCDA flat on the graphene surface. Moreover, STM images measured
films deposited on sulfur-passivated GaAs(001) surfaces.[26] STM at several sample bias allowed the identification of submolecular
results indicate a Stranski–Krastanov growth mode for the RT features. These features are directly related to the PTCDA free
deposition of 14 ML PTCDA. The molecular planes of PTCDA molecule frontier orbitals.

Figure 11. 4.5  4.5 nm2 STM images with intramolecular resolution of PTCDA on graphene/Pt(111). The tunneling conditions are V ¼ 2.3 V,
I ¼ 0.23 nA for (a) and V ¼ þ1.24 V, I ¼ 0.52 nA for (b). The insets compare zoom-in of single molecules from the experimental images with DFT
calculated molecular orbitals corresponding to HOMO (left) and LUMO (right) of the free molecule. Reproduced with permission.[41] Copyright 2014,
American Chemical Society.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (9 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

Figure 12. a) STM image 60  60 nm2 showing an 80 nm long Fe-PTCDA chain. I ¼ 0.21 nA, V ¼ 1.8V. b) Model for the chain-like metal-organic structure.

Figure 13. a) STM image (65  65 nm2) obtained at 390 K. V ¼ 1.5 V; I ¼ 0.23 nA of a Fe-PTCDA monolayer. Several domains (here marked as A–D)
present the “extended ladder-like” structure. b) STM image of a detail of the 2D extended structure (14.2  14.2 nm2). V ¼ 1.0 V; I ¼ 0.4 nA. c) Unit cell
of the DFT calculation of the 2D structure. The electron charge values for the main atoms are indicated. d) DFT calculated molecular orbitals HOMO
(blue–cyan) and LUMO (orange–yellow) for the 2D extended structure showing the different location of these orbitals. Reproduced with permission.[52]
Copyright 2010, IOP Publishing.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (10 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

As observed in Figure 11, the HOMO and LUMO orbitals of of STM/STS shows a wide diversity of results indicating a
the original molecule can be easily recognized on STM images complex behavior of the PTCDA molecules. It has been found
measured at negative and positive sample polarities, V ¼ 2.3 V out that the surface reactivity highly influences the PTCDA
and V ¼ þ1.24 V respectively. This is a clear indication of the adsorption and growth properties. The degree of the molecule–
weak interaction between graphene and PTCDA. This was surface interaction can lead to randomly chemisorbed molecules
further confirmed by first-principles DFTcalculations both of the on highly reactive substrates, to one-dimensional row formation
free molecule and the complete PTCDA/graphene/Pt(111) on moderately reactive surfaces and well-ordered overlayers in
system.[41] the case of physisorbed PTCDA molecules on weakly reactive
As shown in this study, the growth and electronic structure of surfaces.
PTCDA on the weakly coupled graphene on Pt(111) surface
retain all of the essential electronic features of the molecular
layer upon adsorption. Acknowledgements
We thank and congratulate Professor Dietrich R. T. Zahn for coordinating
DIODE project. We also thank all the people participating in the DIODE
6. PTCDA in Metalorganic Nanostructures project, and Luis Rodríguez for critical reading. Funding for this work has
PTCDA overlayers are characterized by a perpendicular been provided by European DIODE-Network by EU Network DIODE
(Contract No HPRN-CT-1999-00164). We acknowledge the financial
disposition[11–14] of contiguous molecules stabilized through
support from AEI and FEDER under projects MAT2017-85089-C2-1-R and
hydrogen bonds between the oxygen atoms of a molecule with MAT2016-77852-C2-2-R (AEI/FEDER, UE) and from MINECO under
the hydrogen atom of the neighbor one. Coadsorption of project MDM-2014-0377 and RTC-2015-3662-2.
transition metals (Fe, Co, Ni) with PTCDA leads to the
formation of metal-organic structures.[51,52] In these, coordina-
tion bonds between iron and oxygen atoms link adjacent
Conflict of Interest
molecules developing chain structures. In Figure 12a a STM
image shows a metal-organic chain formed on Au(111) by The authors declare no conflict of interest.
coadsoption of iron and PTCDA. The structure is pinned at the
domain boundary of the gold reconstruction joining the elbows
of the reconstruction. Such points are preferential adsorption Keywords
places for the iron. The model for this structure is shown in
PTCDA, scanning tunneling microscopy, scanning tunneling
Figure 12b, where two adjacent molecules are linked via two
spectroscopy
iron atoms.[50]
Other metal-organic structures can be formed for higher
coverage.[52] In that case, extra PTCDA molecules placed Received: July 6, 2018
Revised: December 5, 2018
perpendicularly link parallel chains building ladder-like struc-
Published online: January 28, 2019
tures that grow covering the substrate. Figure 13 shows the
structures obtained after the co-evaporation of 0.9 ML of PTCDA
and 0.07 ML of iron and then annealing the surface at 390 K. As
[1] S. R. Forrest, Chem. Rev. 1997, 97, 1793.
result, most of the surface is then covered with a 2D extended
[2] C. Siedel, C. Awater, X. D. Liu, L. Ellerbrake, H. Fuchs, Surf. Sci. 1997,
ladder metal-organic structure. The model is shown in
371, 123.
Figure 13c. This configuration contains nonequivalent PTCDA [3] B. Krause, A. C. Dürr, K. Ritley, F. Schreiber, H. Dosch, D. Smilgies,
molecules, as the ones forming the long 1D chain interact with Phys. Rev. B 2002, 66, 235404.
four iron atoms, while the ones at rung positions have [4] L. Chkoda, M. Schneider, V. Shklover, L. Kilian, M. Sokolowski,
interaction with two iron atoms. This can be observed in C. Heske, E. Umbach, Chem. Phys. Lett. 2003, 371, 548.
STM images at both polarities,[52] where one or the other PTCDA [5] L. Kilian, E. Umbach, M. Sokolowski, Surf. Sci. 2004, 573, 359.
molecules appear brighter respectively. Figure 13c and d show [6] D. Braun, A. Schirmeisen, H. Fuchs, Surf. Sci. 2005, 75, 3.
DFT calculations performed for this structure. The molecular [7] M. Stöhr, M. Gabriel, R. Möller, Surf. Sci. 2002, 330, 507.
orbitals calculated are presented in Figure 13d. In blue colors the [8] Th. Wagner, A. Bannani, C. Bobisch, H. Karacuban, M. Stöhr,
M. Gabriel, R. Möller, Org. Electron. 2004, 5, 35.
calculated HOMO is placed over the chains, and in orange-
[9] P. Fenter, P. E. Burrows, P. Eisenberger, S. R. Forrest, J. Cryst. Growth
yellow the calculated LUMO is placed over the perpendicular
1995, 152, 65.
molecules. Both orbitals resemble to the LUMO of the free [10] P. Fenter, P. Eisenberger, P. Burrows, S. R. Forrest, K. S. Liang, Physica
standing PTCDA molecule, as result of the charge transfer B 1996, 221, 145.
mediated by the iron.[52] [11] T. Schmitz-Hübsch, T. Fritz, R. Sellam, R. Staub, K. Leo, Phys. Rev. B
1997, 55, 7972.
[12] P. Fenter, F. Schreiber, L. Zhou, P. Eisenberger, S. R. Forrest, Phys.
7. Conclusions Rev. B 1997, 56, 3046.
[13] I. Chizhov, A. Kahn, G. Scoles, J. Cryst. Growth 2000, 208, 449.
In this review we present the application of the scanning probe [14] S. Mannsfeld, M. Toerker, T. Schmitz-Hübsch, F. Sellam, T. Fritz,
microscopy and spectroscopy to the study of the structural and K. Leo, Org. Electron. 2001, 2, 121.
electronic properties of organic/inorganic systems. The investi- [15] N. Nicoara, E. Román, J. M. G omez-Rodríguez, J. A. Martín-Gago,
gation of PTCDA on semiconductor or metals surfaces by means J. Mendez, Org. Electron. 2006, 7, 287.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (11 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.pss-b.com

[16] F. Forster, G. Nicolay, F. Reinert, D. Ehm, S. Schmidt, S. Hüfner, Surf. [34] F. Vaurette, J. P. Nys, B. Grandidier, C. Priester, D. Stievenard, Phys.
Sci. 2003, 160, 532. Rev. B 2007, 75, 1.
[17] N. Nicoara, J. Mendez, J. M. Gómez-Rodríguez, Nanotechnology [35] N. Nicoara, J. Mendez, J. M. G omez-Rodríguez, Nanotechnology
2016, 27, 475707. 2016, 27, 365706.
[18] J. Ziroff, P. Gold, A. Bendounan, F. Forster, F. Reinert, Surf. Sci. 2009, [36] E. Ganz, H. Ing-Shouh, X. Fulin, S. K. Theiss, J. Golovchenko, Surf.
603, 354. Sci. 1991, 257, 259.
[19] N. Nicoara, O. Paz, J. Mendez, A. M. Baro, J. M. Soler, J. M. Gomez- [37] M. Hupalo, T. L. Chan, C. Z. Wang, K. M. Ho, M. C. Tringides, Phys.
Rodriguez, Phys. Rev. B 2010, 82, 075402. Rev. B 2002, 66, 1.
[20] K. Iwata, S. Yamazaki, P. Mutombo, P. Hapala, M. Ondracek, [38] S. Stepanovsky, M. Yakes, V. Yeh, M. Hupalo, M. C. Tringides, Surf.
P. Jelinek, Y. Sugimoto, Nature Commun. 2015, 6, 7766. Sci. 2006, 600, 1417.
[21] J. B. Gustafsson, E. Moons, S. M. Widstrand, L. S. O. Johansson, [39] F. S. Tautz, Prog. Surf. Sci. 2007, 82, 479.
Surf. Sci. 2004, 572, 23. [40] Q. H. Wang, M. C. Hersam, Nature Chem. 2009, 1, 206.
[22] T. Soubiron, F. Vaurette, J. P. Nys, B. Grandidier, X. Wallart, [41] A. J. Martinez-Galera, N. Nicoara, J. I. Martínez, Y. J. Dappe,
D. Stievenard, Surf. Sci. 2005, 581, 178. J. Ortega, J. M. G omez-Rodríguez, J. Phys. Chem. C 2014,
[23] A. J. Martinez-Galera, Z. Wei, N. Nicoara, I. Brihuega, J. M. Gomez- 118, 24.
Rodriguez, Nanotechnology 2017, 28, 095703. [42] N. Nicoara, Z. Wei, J. M. G omez-Rodríguez, J. Phys. Chem. C 2009,
[24] E. Umbach, M. Sokolowski, R. Fink, Appl. Phys. A: Mater. Sci. Process. 113, 33.
1996, 63, 565. [43] M. Upward, P. Moriarty, P. Beton, Phys. Rev. B 1997, 56, R1704.
[25] P. Kocan, Y. Yoshimoto, K. Yagyu, H. Tochihara, T. Suzuki, J. Phys. [44] J. Teng, K. Wu, J. Guo, E. Wang, Surf. Sci. 2008, 602, 3510.
Chem. C 2017, 121, 3320. [45] J. C. Swarbrick, J. Ma, J. A. Theobald, N. S. Oxtoby, J. N. O’Shea,
[26] N. Nicoara, O. Custance, D. Granados, J. M. Garcia, J. M. Gomez- N. R. Champness, P. H. Beton, J. Phys. Chem. B 2005, 109,
Rodriguez, A. M. Baro, J. Mendez, J. Phys.: Condens. Matter 2003, 15, 12167.
S2619. [46] J. B. Gustafsson, H. M. Zhang, L. S. O. Johansson, Phys. Rev. B 2007,
[27] G. Goryl, S. Godlewski, J. J. Kolodziej, M. Szymonski, Nanotechnology 75, 1.
2008, 19, 185708. [47] T. G. Gopakumar, J. Meiss, D. Pouladsaz, M. Hietschold, J. Phys.
[28] D. Toton, S. Godlewski, G. Goryl, J. J. Kolodziej, L. Kantorovich, Chem. C 2008, 112, 2529.
M. Szymonski, Phys. Rev. B 2011, 83, 235431. [48] H. T. Zhou, J. H. Mao, G. Li, Y. L. Wang, X. L. Feng, S. X. Du,
[29] O. Paz, I. Brihuega, J. M. Gomez-Rodriguez, J. M. Soler, Phys. Rev. K. Mullen, H. J. Gao, Appl. Phys. Lett. 2011, 99, 153101.
Lett. 2005, 94, 056103. [49] X. Liu, J. Gu, K. Ding, D. J. Fan, X. E. Hu, Y. W. Tseng, Y. H. Lee,
[30] Y. Hirose, S. R. Forrest, A. Kahn, Phys. Rev. B 1995, 52, 14040. V. Menon, S. R. Forrest, Nano Lett. 2017, 17, 3176.
[31] D. Tenne, S. Park, T. Kampen, A. Das, R. Scholz, D. Zahn, Phys. Rev. B [50] R. Forker, T. Dienel, A. Krause, M. Gruenewald, M. Meissner,
2000, 61, 14564. T. Kirchhuebel, O. Groning, T. Fritz, Phys. Rev. B 2016, 93, 165426.
[32] Q. Chen, T. Rada, T. Bitzer, N. V. Richardson, Surf. Sci. 2003, 547, 385. [51] J. Mendez, R. Caillard, G. Otero, N. Nicoara, J. A. Martín-Gago, Adv.
[33] G. Sazaki, T. Fujino, J. T. Sadowski, N. Usami, T. Ujihara, K. Fujiwara, Mater. 2006, 18, 2048.
Y. Takahashi, E. Matsubara, T. Sakurai, K. Nakajima, J. Cryst. Growth [52] L. Alvarez, S. Peláez, R. Caillard, P. A. Serena, J. A. Martín-Gago,
2004, 262, 196. J. Mendez, Nanotechnology 2010, 21, 305703.

Phys. Status Solidi B 2019, 256, 1800333 1800333 (12 of 12) © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Das könnte Ihnen auch gefallen