Skip to main content
Advanced Science logoLink to Advanced Science
. 2020 Feb 14;7(6):1902315. doi: 10.1002/advs.201902315

Charge‐Transfer‐Controlled Growth of Organic Semiconductor Crystals on Graphene

Nguyen Ngan Nguyen 1, Hyo Chan Lee 1, Min Seok Yoo 1, Eunho Lee 1, Hansol Lee 1, Seon Baek Lee 1, Kilwon Cho 1,
PMCID: PMC7080519  PMID: 32195079

Abstract

Controlling the growth behavior of organic semiconductors (OSCs) is essential because it determines their optoelectronic properties. In order to accomplish this, graphene templates with electronic‐state tunability are used to affect the growth of OSCs by controlling the van der Waals interaction between OSC ad‐molecules and graphene. However, in many graphene‐molecule systems, the charge transfer between an ad‐molecule and a graphene template causes another important interaction. This charge‐transfer‐induced interaction is never considered in the growth scheme of OSCs. Here, the effects of charge transfer on the formation of graphene–OSC heterostructures are investigated, using fullerene (C60) as a model compound. By in situ electrical doping of a graphene template to suppress the charge transfer between C60 ad‐molecules and graphene, the layer‐by‐layer growth of a C60 film on graphene can be achieved. Under this condition, the graphene–C60 interface is free of Fermi‐level pinning; thus, barristors fabricated on the graphene–C60 interface show a nearly ideal Schottky–Mott limit with efficient modulation of the charge‐injection barrier. Moreover, the optimized C60 film exhibits a high field‐effect electron mobility of 2.5 cm2 V−1 s−1. These results provide an efficient route to engineering highly efficient optoelectronic graphene–OSC hybrid material applications.

Keywords: charge transfer, graphene, growth template, organic electronics, organic semiconductors


The growth behavior of fullerene (C60) thin films on graphene templates where charge transfer occurs is presented. The number of electrons transferred from graphene to C60 is controlled by in situ electrical gating of graphene during C60 deposition. When this electron transfer is suppressed, high‐crystalline C60 thin films are achieved for highly efficient optoelectronic applications.

graphic file with name ADVS-7-1902315-g007.jpg

1. Introduction

Graphene has excellent properties, so the possibility of integrating it with both inorganic and organic semiconductors has been intensively studied. Graphene–semiconductor heterostructures provide multifunctionality and desirable properties for scalable and flexible optoelectronic applications.1, 2 The ideally sp2‐hybridized carbon atoms of graphene constitute a basal plane with no dangling bonds, so it provides an atomically clean interface with a semiconductor; this contact is extraordinary and cannot be achieved with traditional interfaces. With the introduction of these unique graphene–semiconductor interfaces, researchers have proposed various graphene–semiconductor hybrid optoelectronic devices such as field‐effect transistors (FETs), light‐emitting diodes, solar cells, photodetectors, and barristors.3, 4, 5

Graphene is inert and is composed of a single‐atom‐thick layer, so it is a useful growth template for semiconductors, especially organic semiconductors (OSCs).6, 7 The assembly of OSC thin films on graphene is mainly determined by the interactions between OSC ad‐molecules and the graphene template (e.g., van der Waals). Therefore, the graphene template can enable epitaxial growth of highly crystalline OSC thin films.8 In addition, these interactions can easily be tuned by controlling the electronic properties of graphene,9, 10 so graphene templates offer a facile and direct approach to prepare graphene–OSC heterostructures with desirable interfacial properties. However, despite the great potential of graphene–OSC heterostructures, only a few studies of OSCs' growth behavior on electronic‐states‐controlled graphene have been reported.7, 10 Therefore, to develop a reliable method to optimize the growth of OSCs on graphene templates, the complex of OSC molecules and graphene templates and possible interactions between them should be investigated.

Here, we demonstrate that an epitaxial growth of a vacuum‐deposited fullerene (C60) thin film on a graphene template can be controlled by tuning charge transfer between them. The Fermi level (E F) of the graphene template determines the amount of charge transfer between the graphene and the C60 ad‐molecules, and this amount in turn affects the molecular dynamics of C60 on the graphene template. By finely tuning the E F of the graphene template, we induced layer‐by‐layer growth of highly ordered C60 films on graphene. Considering that the thin film's topological and crystalline features determine the optoelectronic properties of OSCs,11 this approach advances the efficiency of organic electronic devices. The C60 films grown under optimized conditions exhibited a maximum field‐effect mobility of 2.5 cm2 V−1 s−1. Furthermore, a graphene–C60 Schottky junction prepared by our method approached the Schottky–Mott limit, which is desirable for highly efficient graphene–OSC barristors and other vertical graphene–OSC hybrid optoelectronic devices.

2. Results and Discussion

2.1. Charge Transfer between Graphene and C60

We first investigated the transfer of electrons from graphene to C60. Analyses using ultraviolet photoelectron spectroscopy, Kelvin probe force microscopy, and Raman spectroscopy revealed that the adsorption of C60 molecules induced p‐type doping of graphene (Figure S2, Supporting Information). To clarify the relationship between charge transfer and the initial electronic states of graphene, we fabricated graphene field‐effect transistors (G‐FETs) on 300 nm thick SiO2/Si substrates and compared the transfer characteristics of the G‐FETs before and after 3 s of C60 deposition at a deposition rate of 5 × 10−2 monolayer per second (ML s−1) (Figure 1 a). To eliminate the contact resistance, we used transfer‐length‐method measurements so that the change in graphene channel resistance (R Ch) could be solely attributed to the change in charge‐carrier density (n g, n g > 0 for electrons and n g < 0 for holes).

Figure 1.

Figure 1

Charge transfer between graphene and C60. a) Schematic diagram showing G‐FET with deposited C60. b) Transfer characteristic of G‐FET before (green open circle) and after C60 deposition (blue closed circle). Solid lines are model fits. c) Concentration of transferred charge carrier after C60 deposition Δn CT versus initial charge carrier concentration of bare graphene n g,bare. d) Energy band diagrams of graphene/C60 when n g,bare < n c (left), n g,bare = n c (middle), and n g,bare > n c (right).

After C60 deposition, the R Ch was preserved as long as the gate voltage (V G) was <−40 V. This preservation demonstrates that deposition of C60 did not cause degradation of graphene, and more importantly, that no charge transfer occurred between graphene and C60 in this range of V G. However, at V G > −40 V, the R Ch –V G curve shifted to the right; this change indicates that electrons were transferred from graphene to C60 (Figure 1b). This shift of R Ch –V G curves when the magnitude of V G is larger than a certain value was consistently observed with other samples from different batches (Figure S3, Supporting Information).

To calculate the number of transferred electrons (Δn CT (cm−2)) at a certain V G, the R Ch –V G curve was fitted using the constant‐mobility model.12 Then the carrier density of bare graphene before C60 deposition (n g,bare) and the carrier density of graphene–C60 after C60 deposition (ng,C60) were each calculated at each V G as

ng=sgnVGVD1μeRChLW2nres2 (1)

where V D is V G at maximum R Ch, μ is the carrier mobility, e is the elementary charge, L is the channel length, W is the channel width, and n res is the residual carrier concentration in graphene. Then Δn CT was calculated as ng,bareng,C60. Before C60 deposition, the fitted values of μ and n res of the graphene transistor were 4470 cm2 V−1 s−1 and 2.3 × 1012 cm−2, respectively. When plotted versus n g,bare (Figure 1c), extracted Δn CT showed no charge transfer between graphene and C60 when n g,bare was less than a critical value, n c = −4.4 × 1012 cm−2. As n g,bare approached n c, charge transfer started and gradually increased with increasing n g,bare. The V G‐dependent contact resistance in G‐FETs also supports our claim that the charge transfer occurred when n g,bare > n c (Figure S4, Supporting Information).

The observed n g,bare‐dependent charge transfer between graphene and C60 is explained as follows. The electrons in graphene are transferred to C60 when the E F of graphene is higher than the lowest unoccupied molecular orbital (LUMO) level of adjacent C60. The LUMO level of isolated C60 molecules is known to be −4.5 eV,13 which is similar to the E F of undoped graphene. However, the energy levels of organic molecules change and broaden upon adsorption of C60, because of the polarizability of the substrate;14, 15 thus, the LUMO level of C60 adsorbates can lie below the E F of undoped graphene that has n g,bare > n c; as a result, the graphene becomes p‐type doped. The absence of charge transfer when n g,bare < n c is attributed to the E F of graphene being lower than the LUMO level of the C60 adsorbates (Figure 1d, left). As the E F of graphene is raised by external gating such that it reaches the LUMO level of C60, electrons are transferred from graphene to C60, and the E F of graphene is pinned to the LUMO level of C60. As a result of this charge transfer, an electric field is generated between the graphene and the C60, so the vacuum level at the interface becomes tilted so that the E F of the graphene and the LUMO level of the C60 are aligned (Figure 1d, right).

First, the number of charges is conserved at the graphene–C60 interface as

CgeVGVD,bare=ng,bare=ng,C60+σC60e (2)

where C g is the dielectric capacitance, V D,bare is the V D of the G‐FET before C60 deposition, and σC60 is the surface charge density in a C60 film. The charge redistribution at the graphene–C60 interface as a function of n g,bare can be estimated by solving

sgnng,C60vFπ|ng,C60|sgnncvFπ|nc|=eσC60ε0d (3)

where ℏ is the reduced Planck's constant, v F is the Fermi velocity of graphene, ε0 is the vacuum permittivity, and d is a fitting parameter that describes the spacing between graphene and C60. The left‐hand side of Equation (3) is E FE F,c (Figure 1d), in which E F,c is the critical Fermi level where the charge transfer between graphene and C60 occurs. The right‐hand side is the charge‐transfer‐induced shift of the vacuum level at the interface.

The R ChV G curves of G‐FETs and the Δn CT as a function of n g,bare were modelled using calculated ng,C60 and d. The models successfully replicated the experimental values (Figure 1b,c). Moreover, the charge transfer modifies the density of states of C60 so that the LUMO level of charged C60 molecules is split into an “occupied” LUMO level (L1) that is shifted downward and an unoccupied LUMO level (L2) that is shifted upward (Figure 1d, right).16 This downshift of the LUMO level upon charge transfer can substantially stabilize C60 adsorbates on graphene.17

2.2. Growth of C60 Thin Films on Graphene under Charge Transfer

With the in situ electrical gating of graphene (“Experimental and Methods” in the Supporting Information), we observed changes in i) the molecular interactions and assembly of C60 ad‐molecules and ii) the growth behavior of C60 crystals as the E F of graphene gradually approached the E F,c.

First, C60 ad‐molecules may interact with each other on the graphene surface, depending on the relative position of the E F of graphene and the E F,c. These distinctions can be well detected by Raman spectroscopy (Figure S5, Supporting Information). In both Raman spectra, the feature peaks of C60, i.e., the A g(1) mode at ≈500 cm−1 and the A g(2) mode at ≈1470 cm−1, were clearly observed.18 The position of the A g(2) peak indicates the number of intermolecular bonds to each C60 molecule, where each intermolecular bond shifts the peak by −5 cm−1.19 The peak position of the A g(2) mode of C60 grown on graphene with E F < E F,c is consistent with that reported for pristine C60 molecules.18, 19 However, the A g(2) peak of C60 grown on graphene with E F > E F,c was red‐shifted ≈3 cm−1; this change indicates that chemically bonded C60 dimers or oligomers were formed. This selective formation at high E F strongly suggests that control of the E F of graphene during C60 growth indeed determined the charge state of the C60 ad‐molecules.

The charge state of C60 ad‐molecules determines the formation of covalent bonds between two C60 molecules.20, 21 When C60 molecules have negative charges, the activation barrier for the bonding decreases. Therefore, graphene with E F > E F,c induced negative charges in C60 ad‐molecules, resulting in the formation of intermolecular bonds between C60 ad‐molecules. By contrast, on graphene with E F < E F,c, C60 molecules were charge‐neutral and thus did not form covalently bonded C60 dimers.

Δn CT affected molecular arrangement in C60 crystals, and consequently, how those crystals assembled into thin films. We used grazing incidence X‐ray diffraction (GIXD) to characterize C60 thin films with different thicknesses grown on graphene, where Δn CT was controlled. Under ambient conditions, the most stable structure of C60 crystals is face‐centered cubic (fcc);22 the diffraction patterns of the fcc C60 were observed in our system of C60 thin films grown on graphene (Figure 2 a).

Figure 2.

Figure 2

Crystal structure of C60 films grown on graphene. a) 2D GIXD patterns of 2.5 ML (2 nm) and thick (100 nm) C60 films grown on graphene when Δn CT = 0 cm−2 (left), Δn CT = 5 × 1011 cm−2 (middle), and Δn CT = 1.3 × 1012 cm−2 (right) during C60 deposition. b) Cross‐sectional profiles of the 2D GIXD image along the qz for various Δn CT. c) The mean size of the crystalline (111) domains R (111) versus Δn CT. d) Schematic illustrations of C60 crystal growth on graphene without (upper) and with (lower) the charge transfer between them. Insets: Low‐magnification HR‐TEM images of corresponding graphene–C60 samples on TEM grids. Scale bar in insets: 200 nm.

At the early growth stage (nominal thickness of 2.5 ML), irrespective of the occurrence of charge transfer, the set of reflections of (111) family and the reflections of plane (113) and plane (220) appeared; these reflections are located along the out‐of‐plane direction (qz) and at 30° and 35° tilt from qz, respectively. These results indicate that C60 has an epitaxial relationship with graphene, with the (111) plane of C60 crystals parallel to the graphene substrate;23 this epitaxy was independent of Δn CT. However, differences were observed in the crystal domain sizes of C60 thin films grown on graphene at different Δn CT (Figure 2c). We quantified the average crystal domain size of C60 thin films by using the Scherrer equation to estimate the domain sizes of crystal plane (111) (R (111)). When Δn CT = 0 during C60 growth, C60 thin films had R (111) ≈ 60 nm, which is almost three times larger than in the film grown under very high Δn CT.

At the final growth stage, the GIXD patterns of C60 films grown with and without charge transfer both showed clear ring patterns, which reveal the presence of randomly oriented C60 crystals. However, the thick C60 films' ordering degree was still strongly dependent on Δn CT. On the graphene surface where Δn CT = 0, the reflections were still sharp with a high signal‐to‐noise ratio, i.e., most of the C60 crystals were oriented. As Δn CT increased, these reflections weakened and eventually became undetectable; this change suggests that a large fraction of newly nucleated C60 crystals were randomly oriented on the pre‐existing C60 thin film. The growth behavior of C60 crystals on graphene, as indicated by GIXD experiments, is summarized as follows (Figure 2d). A highly crystalline film of fcc C60 was epitaxially formed on graphene via a layer‐by‐layer growth mode at negligible Δn CT during C60 growth. When Δn CT > 0, despite the epitaxial relationship between graphene and C60 at the early growth stage, randomly oriented nucleation occurred during vertical growth. These inferences are confirmed by low‐magnification high‐resolution transmission electron microscopy (HR‐TEM) images (insets of Figure 2d). At Δn CT = 0, large‐area C60 layers were observed; by contrast, at very high Δn CT, small C60 clusters formed. Although the GIXD results provided a hint about the crystal structure of the C60 films grown on graphene over a macro area, they could not directly reveal the arrangement among C60 molecules and the carbon atoms in graphene.

Therefore, C60 thin films (2.5 ML) grown on graphene were imaged at high magnification using HR‐TEM. The image of C60 grown on graphene at Δn CT = 0 clearly showed an ordered hexagonal arrangement of C60 molecules over a few tens of nanometers, which is the fashion of the (111) plane of a highly crystalline fcc structure (Figure 3 a, top). Moreover, the ordering in this HR‐TEM image matches that of ABA‐stacked C60 layers.24 This stacking order was uniform over the analyzed areas; this consistent order implies that C60 layers were preferentially stacked on each other in an ABA manner when the thin film was grown on graphene at Δn CT = 0. The corresponding selected‐area electron diffraction (SAED) pattern of this C60 thin film also showed only a single set of hexagonal patterns, i.e., the crystalline orientation of C60 was uniform along the vertical direction. Notably, when Δn CT = 0 was maintained during C60 growth, the misorientation angles between the SAED patterns of C60 and those of graphene were concentrated at close to 0° and 30°, which correspond to energetically stable adsorption sites of C60 molecules along the armchair and zigzag directions of graphene, respectively (Figure S7f, Supporting Information).23 This result is further evidence of an epitaxial relationship between graphene and C60.

Figure 3.

Figure 3

Epitaxial molecular arrangement of C60 on graphene. a) Typical HR‐TEM images and SAED patterns of 2.5 ML C60 grown on graphene when Δn CT = 0 (top) and Δn CT >> 0 (bottom). Scale bars in HR‐TEM images: 3 nm; in SAED patterns: 1 nm−1. Insets: High‐magnification HR‐TEM images of regions with ABA (in (top)) and ABC (in (bottom)) stacking. b) Histogram plots of nearest neighbor C60–C60 molecule distances extracted from HR‐TEM images when the growth associated without (left) and with (right) charge transfer. c) DFT energetic simulations of C60–C60 double‐bonded dimer (left) and isolated C60 molecules (right).

By contrast, when C60 was grown on graphene under a very high Δn CT, HR‐TEM image and the corresponding SAED patterns (Figure 3a, bottom) typically revealed polycrystalline C60 thin film along the lateral direction and vertical direction. This C60 film showed ABA and ABC stacking mixed within small areas. In addition, small crystalline domains were tilted from the rest with a high angle (≈30°) in this film (Figure 3a, bottom left). Notably, the areas between the tilt grains mostly exhibited an amorphous structure. On top of this amorphous region, C60 molecules could not arrange well, so the results were i) randomly oriented nucleation of C60 crystals and ii) the formation of additional amorphous layers, or both. The resulting richness of tilt grain boundaries could result in the observed polycrystallinity along both the lateral and vertical directions. The dominance of (111)‐plane‐oriented C60 crystal domains (Figure 2a) suggests the presence of an epitaxial relationship between C60 and graphene at this small thickness, so grains that have high tilt angle may be formed by stitching C60 domains aligned along the armchair direction and those aligned along the zigzag direction of graphene.

HR‐TEM was also the best tool to investigate the chemically bonded dimers in C60 films (Figure S5, Supporting Information). To quantize the dimer content, we analyzed numerous intermolecular distances of two nearest‐neighbor C60 molecules in films grown at Δn CT = 0 and Δn CT > 0 (Figure 3b). In both cases, the distance distribution showed peaks centered near 0.86 and 0.95 nm, but the relative peak heights depended on Δn CT. We could assign the 0.85 nm peak to double‐bonded C60 dimers, and the 0.95 nm peak to isolated C60 molecules.25 When Δn CT = 0, more than half of the intermolecular distances were close to 0.95 nm; this consistent separation implies that a large portion of the C60 molecules were still free and intact. However, at Δn CT > 0 the fraction of free C60 molecules was substantially reduced and the proportion of double‐bonded dimers increased. These results qualitatively show that charge transfer with graphene during C60 growth promoted the formation of double‐bonded C60 dimers.

In addition, we performed density functional theory (DFT) simulations to calculate the electronic structure of a double‐bonded C60 dimer and two isolated C60 molecules (Figure 3c; Figure S14, Supporting Information). Compared with isolated C60 molecules, a double‐bonded C60 dimer showed an ≈0.2 eV smaller bandgap, and broader LUMO and highest occupied molecular orbital (HOMO) levels.

The effects of charge transfer on C60 growth behaviors are further demonstrated by morphological analysis using atomic force microscopy (AFM) (Figure 4 a), which enabled statistical analysis of average height h i of C60 islands and surface coverage θ of the thin films during the early growth stage (Figure 4b). On the surface of graphene templates on which charge transfer was suppressed, i.e., E F < E F,c, the initial large‐area C60 islands expanded laterally, to yield a constant monolayer thickness (0.8 nm) and a large increase of surface coverage. As electron transfer from the graphene to C60 ad‐molecules increased, the number of nuclei quickly increased and each of them merely grew in height; the result was an array of grains of different heights. At Δn CT = 0, as the growth continued, continuous C60 film was formed by coalescence of large‐area C60 grains; by contrast, at Δn CT > 0, C60 film was formed by full coverage of small C60 islands with poor inter‐grain connection. At the later growth stage (12.5 ML), the C60 thin film grown at Δn CT = 0 revealed clear terrace structure, which is evidence of lateral growth mode, whereas the film grown at Δn CT > 0 simply showed an array of tiny crystallites.

Figure 4.

Figure 4

Nucleation of C60 islands on graphene. a) AFM images of C60 at different nominal thicknesses of 0.25 ML (left), 1.25 ML (middle), and 12.5 ML (right) grown on graphene, without (upper) and with (lower) the charge transfer. Scale bar: 400 nm. b) Height analysis for C60 islands in the AFM images. Inset: Surface coverage analysis. c) Nucleation density N i versus Δn CT from gate‐bias (green circle) and polymer‐contact doping (blue square). d) N i versus thermal parameter 1/(k B T). e) Nucleation energy barrier of C60 (E Nuc) versus Δn CT calculated from (d). Shaded areas are to guide the eye.

The charge transfer in the graphene–C60 system as well as its effects on the crystal structure and morphology of C60 (Figures 2, 3, 4) were elucidated using electrically gated graphene templates. The use of polymer–substrate‐doped graphene revealed similar results (Figures S1, S6, and S9, Supporting Information). This comparison emphasizes that other factors (e.g., localized traps, the wetting transparency, or contamination on the graphene surface) that might obscure the collected results might have been effectively eliminated.7 Moreover, this polymer–substrate doping method could provide a general understanding of the observed phenomena.

To quantify the dependence of C60 growth on the charge transfer from the graphene template to C60 ad‐molecules, numerous C60 thin films with a nominal thickness of 0.25 ML were grown on graphene templates whose E F was finely controlled by either gating or polymer–substrate doping. The plot of the nucleation density (N i) of these films against Δn CT at room temperature revealed correlations between the nucleation of C60 and charge transfer from graphene to C60 (Figure 4c; Figure S10, Supporting Information). Clearly, N i increased as Δn CT increased. We also directly measured the activation energy for C60 nucleation (E Nuc) as a function of Δn CT, as N i = C exp (E Nuc/(k B T)) where C is a pre‐exponential factor, k B is the Boltzmann constant, and T is the substrate temperature.26 To this end, N i values as a function of the substrate temperature T were collected at various fixed Δn CT; the slopes of plots of ln(N i) versus 1/(k B T) at a certain Δn CT gave the values of E Nuc at the Δn CT (Figure 4d). As a result, we confirmed that E Nuc increased as Δn CT increased (Figure 4e).

2.3. Atomistic Mechanism of C60 Thin Film Growth on Graphene under Charge Transfer

The nucleation of C60 on graphene involves several atomistic processes (Figure 5 a). After adsorbing to graphene, an C60 ad‐molecule diffuses on the surface until the molecule forms a dimer with another ad‐molecule or attaches to a pre‐existing island (growth).27, 28 In general, E Nuc is related to the activation energies of all of these atomistic processes. However, the energy barrier for C60 diffusion is negligible on graphitic surfaces,24, 29 so nucleation and growth of C60 on graphene are predominantly limited by the rate of attachment of ad‐molecules to pre‐existing islands.

Figure 5.

Figure 5

Mechanism of C60 growth on graphene. a) Nucleation process of C60 crystals on graphene surface that involves adsorption, diffusion, dimer formation, attachment, and direct impingement. b) Energy profiles of a C60 ad‐molecule versus position near and on a C60 island under the absence (solid line) and presence (dashed line) of the charge transfer between the ad‐molecule and graphene.

For such attachment‐limited nucleation with negligible barriers to diffusion and dimerization, E Nuc = [2E i + 2(i + 1)E B]/(i + 3),27 where i is critical cluster size, E i is cluster energy, and E B is activation energy for the attachment.27 This equation implies that a nucleation density increases as E B increases. This relation is explained as follows. The presence of high E B hinders the attachment of deposited ad‐molecules to an island, so the concentration of ad‐molecules increases on the graphene surface. Thus, the probability of ad‐molecules colliding rapidly increases, and this change favors new nucleation rather than the growth of pre‐existing islands.

Therefore, the increases in N i and E Nuc with increasing Δn CT are attributable to the increase in E B with increasing Δn CT as E B (Δn CT) = E B0  + EBn CT) where E B0 is the charge‐transfer‐independent attachment barrier and EB is the charge‐transfer‐dependent attachment barrier. When electrons in graphene are transferred to the ad‐molecules and the islands, the ad‐molecules and islands are negatively charged and the underlying graphene becomes positively charged (Figure 5b; Figure S11, Supporting Information). Consequently, repulsive Coulomb interaction occurs between the dipole from the ad‐molecule–graphene and that from the island–graphene. This long‐range repulsive interaction would introduce an additional attachment barrier EB. Assuming the long‐range repulsive interaction is simple electrostatic repulsive interaction, EB can be estimated as

EB=Zavge2dΔnCT/2ε0 (4)

where Z avg is the average charge state of C60 ad‐molecules (Equation (4) is derived in the Supporting Information). This model successfully predicts the increase in E B with increasing Δn CT. In this argument, we assumed that repulsive Coulomb interaction between an ad‐molecule and an island (and not that between two ad‐molecules on graphene) dominantly affects the nucleation kinetics. This assumption can be justified because the probability of collision between two C60 molecules which both simultaneously have negative charges would be very small. On the contrary, a C60 island contains many C60 molecules, so a C60 island is likely negatively charged.

The transition from a 2D to a 3D growth mode (Figures 2, 3, 4) under the charge transfer between graphene and C60 can be simply explained by invoking the repulsive Coulomb interaction between an ad‐molecule and an existing island. With increasing Δn CT, E B increases because of the repulsive interaction; this change inhibits the lateral growth of negatively charged islands by negatively charged ad‐molecules diffusing on the graphene surface. However, irrespective of the E F of the graphene template, the ad‐molecules from the vapor phase can land directly on the top of the existing island because they are charge‐neutral and thus not prone to the repulsive Coulomb interaction. However, after they are deposited on the top of the islands, their dynamics are again influenced by the E F of graphene. When Δn CT = 0, they can move relatively freely down to the graphene surface because the Ehrlich–Schwoebel barrier is much lower than the diffusion barrier on top of the C60 layer.28 When electrons are transferred from graphene to C60, the E B increases and thus acts as an energy wall surrounding the edge of islands. For an ad‐molecules on the top of the island to move downward and escape from the island, they must overcome an activation energy greater than E B. Consequently, ad‐molecules become concentrated on the top of the island, so the island rapidly grows in the vertical direction. The rapid vertical growth in turn leads to the formation of randomly oriented crystals.

2.4. Charge Transport in C60 Thin Films and Graphene–C60 Junctions

To quantify the advantage of our growth‐controlled C60 thin films for lateral charge transport, we grew C60 thin films on graphene at controlled charge‐transfer conditions, then transferred the C60 films to octadecyltrichlorosilane (ODTS)‐treated SiO2/Si substrates and then fabricated planar C60 transistors (C60‐FETs). The final device included an ≈50 nm thick C60 channel without the underlying graphene (Figure 6 a). We measured the transfer characteristics of C60‐FETs in the saturation regime with C60 thin‐film channels grown at different Δn CT, then estimated the associated electron field‐effect mobility (μe) and measured the on/off ratio (I on/I off). For a C60 thin film grown at Δn CT = 0, the I on/I off of the FET device was ≈107 and the average μe was ≈1.5 cm2 V−1 s−1. The maximum mobility of the device was ≈2.5 cm2 V−1 s−1, which is similar to the state‐of‐the art mobility of C60 transistors fabricated by the vapor deposition method (Figure 6b).5, 30 With increasing Δn CT, the I on/I off and μe of the device substantially decreased, and eventually reached the same level of devices fabricated with polycrystalline and small‐grain C60 (Figure 6c).31 The decay occurs because the high Δn CT causes low crystallinity, low uniformity and limited grain size, and these traits suppress the lateral μe of C60 thin films.

Figure 6.

Figure 6

C60 field‐effect transistors and graphene–C60 barristors. a) Schematic illustration of planar C60‐FET. b) Transfer characteristic of C60‐FET with C60 film grown at Δn CT = 0. c) Average I on/I off and electron mobilities μe of C60‐FETs versus Δn CT during C60 growth. d) Schematic illustration of graphene–C60 barristor. e) I DS versus V DS of graphene–C60 barristors at various fixed V G (from −100 to −40 V, step 10 V) for Δn CT = 0 (left) and at V G (from −100 to 100 V, step 10 V) for Δn CT > 0 (right). Inset: I DS versus V DS at linear scale of graphene–C60 barristor for Δn CT = 0 at V G = −40 V (filled symbols) and V G = −30 V(open symbols). f) Temperature‐dependent saturation current of graphene–C60 barristors at various V G for Δn CT = 0 (left, step 10 V) and Δn CT > 0 cases (right, step 40 V). g) The Schottky barrier height (ΦB) obtained from (f) versus ΔE F.

Our method of growing C60 thin films on graphene provides a direct way to produce controlled graphene–C60 van der Waals heterostructures. In addition to its use as a growth template, graphene can function as an active layer or electrode for various flexible optoelectronic devices because of its excellent electrical conductivity and flexibility. Recently, heterostructures composed of graphene and OSCs have shown promise for use in organic photovoltaics, organic light‐emitting diodes, organic photodetectors, and vertical FETs.4, 5 The electrical characteristics of such devices depend on the charge‐injection efficiency at the graphene–OSC interface.

Such charge‐injection efficiency at the graphene–C60 van der Waals heterointerface formed with our method was demonstrated by fabricating two types of graphene–C60 barristors. They had the same device structure, but one had C60 film grown at Δn CT = 0, and one had C60 film grown at Δn CT = 1 × 1012 cm−2 (Figure 6d), so the C60 layers had enormously different morphological and crystalline features. Both devices showed typical n‐type barristor behavior (Figure 6e).1 The closeness between the LUMO level of C60 and Fermi level of aluminum yields Ohmic contact between the C60 and the top aluminum electrode,32 so rectifications arose from the Schottky barrier (ΦB) between the C60 layer and the bottom graphene. Increasing the V G barely affected the current in the forward regime (V DS < 0) but boosted the current in the reverse regime (V DS > 0). Consistent with the band diagram (Figure 1), the increase of V G raised the Fermi level of graphene closer to the LUMO level of C60, reducing the ΦB accordingly, until alignment was achieved between them (ΦB ≈ 0, Ohmic contact).

Although both barristors showed rectification behavior, great distinction was observed in the current levels between the two devices. The device that used the C60 layer that had been grown at Δn CT = 0 showed substantial modulation of the reversed current by the gate voltage; and the on‐state current I on was higher in this device than in the device that used the C60 layer that had been grown at Δn CT > 0, whereas their off‐state currents I off were similar. As a result, this device fabricated with a highly crystalline C60 film (i.e., grown at Δn CT = 0) achieved an I on/I off ratio of ≈103 at V DS = 2 V, which is nearly two orders of magnitude greater than the I on/I off ratio of the other device at the same V DS (Figure 6e; Figure S12, Supporting Information).

The most important difference between the two types of barristors was the occurrence of a Schottky‐to‐Ohmic transition, which was only observed in the device that used the C60 layer that had been grown at Δn CT = 0 (Figure 6e, left). This transition occurred at V G = −40 V, which is consistent with the critical voltage (V G at n c) required to induce charge transfer between graphene and C60 (Figure 1). By contrast, Schottky‐to‐Ohmic transition was not observed within the wider examining V G range for the device with the C60 layer grown at Δn CT > 0 (Figure 6e, right); this absence implies that modulation of the E F of graphene by electrical gating was limited at the graphene–C60 interface.

Fermi‐level pinning can occur when there are interfacial states in the HOMO–LUMO gap of C60 layers near graphene.33 To quantitatively analyze the Fermi‐level pinning, we used the diode equation in the reverse bias saturation regime, IDST2expeΦBkBT.1 The value of ΦB at each V G was then estimated from the plot of ln(I DS/T 2) versus 1/(k B T) (Figure 6f). ΦB increased with increasing E F at different rates in the two device types (Figure 6g). For the barristor with the C60 layer grown at Δn CT = 0, the slope S = dΦB/dE F was ≈0.9, which indicates that the graphene–C60 junction in this device approached the Schottky–Mott limit.14, 34 This result demonstrates an atomically clean interface between graphene and the C60 thin film, which has not been previously achieved.23, 35 To achieve this clean heterointerface for the effective tuning of the Schottky barrier, C60 must be deposited directly on a thermally cleaned graphene surface,36 and the electronic state of graphene must be optimized to effectively limit the charge transfer during growth to enable growth of high‐crystallinity C60 film at the interface with graphene. The latter effect of charge transfer during the growth of OSCs has been neglected previously.

For the other device, S was only 0.1, which is indicative of strong Fermi‐level pinning effect at the graphene–C60 interface. The C60 thin film grown on graphene with charge transfer had small and poorly connected C60 grains near the graphene surface (Figures 2, 3, 4), so the interface had i) a high density of C60 grain boundaries, ii) large amorphous areas, and iii) other crystalline defects that would introduce numerous interfacial trap states (Figure S13c, Supporting Information). The Fermi level of graphene was pinned at those states. In addition, because the DFT results reveal a smaller bandgap of a C60 dimer compared with two isolated C60 molecules (Figure 3c), the presence of C60 dimers would introduce shallow charge traps, which can further contribute to the observed Fermi‐level pinning at the graphene–C60 interface.

To directly confirm the interfacial states between the graphene and the C60, photocurrents of G‐FETs fabricated with deposited C60 thin films (20 nm) were measured under light illumination at 0.62 eV (Figure S13a, Supporting Information). For comparison, C60 thin films were grown on top of graphene channels under Δn CT = 0 and Δn CT > 0. At a high positive gate bias (V G = 80 V), only the G‐FET with the C60 thin film grown at Δn CT > 0 showed additional photocurrent as the device was illuminated (Figure S13b, Supporting Information). The excitation energy is much smaller than the bandgap of the C60 thin film and smaller than 2|E F| of graphene at V G = 80 V, so the interband transitions are forbidden in both the C60 thin film and the graphene.37 Therefore, the photocurrent in the G‐FET with a C60 thin film was merely a result of detrapped electrons from the interfacial states, which were abundant in the layer grown at high Δn CT. In fact, we observed positive photocurrent from the G‐FET with the C60 film grown at Δn CT > 0, but observed no photoresponse from the device with C60 film grown at Δn CT = 0.

3. Conclusion

We observed that charge transfer within the graphene–C60 system during the growth of C60 crystals on a graphene template governed such growth and, thus governed the thin film's corresponding crystal structure and morphology. These charge‐transfer phenomena altered the electronic states of the graphene–C60 system, forming negatively charged C60 nuclei and ad‐molecules. Under these conditions, the growth of C60 on graphene was favored in the vertical dimension because of the high attachment barrier energy, resulting thin films with small and randomly oriented crystallites. With this understanding, we proposed that the optimized graphene template for layer‐by‐layer growth of C60 with large and uniformly oriented crystals is the graphene in which the charge transfer from graphene to C60 is suppressed during the C60 growth. Barristors fabricated with this graphene–C60 van der Waals heterostructure showed efficient tunability of the charge injection barrier, approaching the Schottky–Mott limit. In addition, the lateral electron mobility μe in a planar C60‐FET was also boosted to a maximum μe = 2.5 cm2 V−1 s−1.

Conflict of Interest

The authors declare no conflict of interest.

Supporting information

Supporting Information

Acknowledgements

N.N.N. and H.C.L. equally contributed to the work. This work was supported by a grant (Code No. 2011‐0031628) from the Center for Advanced Soft Electronics under the Global Frontier Research Program of the Ministry of Science and ICT, Korea. The authors thank the Pohang Accelerator Laboratory for providing the synchrotron radiation sources at 4D, 3C, and 9A beamlines used in this study.

Nguyen N. N., Lee H. C., Yoo M. S., Lee E., Lee H., Lee S. B., Cho K., Charge‐Transfer‐Controlled Growth of Organic Semiconductor Crystals on Graphene. Adv. Sci. 2020, 7, 1902315 10.1002/advs.201902315

References

  • 1. Yang H., Heo J., Park S., Song H. J., Seo D. H., Byun K. E., Kim P., Yoo I., Chung H. J., Kim K., Science 2012, 336, 1140. [DOI] [PubMed] [Google Scholar]
  • 2.a) Kim K., Choi J.‐Y., Kim T., Cho S.‐H., Chung H.‐J., Nature 2011, 479, 338; [DOI] [PubMed] [Google Scholar]; b) Wang Q. H., Hersam M. C., Nat. Chem. 2009, 1, 206; [DOI] [PubMed] [Google Scholar]; c) Colson J. W., Woll A. R., Mukherjee A., Levendorf M. P., Spitler E. L., Shields V. B., Spencer M. G., Park J., Dichtel W. R., Science 2011, 332, 228; [DOI] [PubMed] [Google Scholar]; d) Di Bartolomeo A., Phys. Rep. 2016, 606, 1. [Google Scholar]
  • 3.a) Lee W. H., Park J., Sim S. H., Lim S., Kim K. S., Hong B. H., Cho K., J. Am. Chem. Soc. 2011, 133, 4447; [DOI] [PubMed] [Google Scholar]; b) Song Y., Li X., Mackin C., Zhang X., Fang W., Palacios T., Zhu H., Kong J., Nano Lett. 2015, 15, 2104. [DOI] [PubMed] [Google Scholar]
  • 4.a) Jo S. B., Kim H. H., Lee H., Kang B., Lee S., Sim M., Kim M., Lee W. H., Cho K., ACS Nano 2015, 9, 8206; [DOI] [PubMed] [Google Scholar]; b) Koppens F. H. L., Mueller T., Avouris P., Ferrari A. C., Vitiello M. S., Polini M., Nat. Nanotechnol. 2014, 9, 780; [DOI] [PubMed] [Google Scholar]; c) Han T.‐H., Lee Y., Choi M.‐R., Woo S.‐H., Bae S.‐H., Hong B. H., Ahn J.‐H., Lee T.‐W., Nat. Photonics 2012, 6, 105. [Google Scholar]
  • 5. Shih C. J., Pfattner R., Chiu Y. C., Liu N., Lei T., Kong D., Kim Y., Chou H. H., Bae W. G., Bao Z., Nano Lett. 2015, 15, 7587. [DOI] [PubMed] [Google Scholar]
  • 6.a) Nguyen N. N., Jo S. B., Lee S. K., Sin D. H., Kang B., Kim H. H., Lee H., Cho K., Nano Lett. 2015, 15, 2474; [DOI] [PubMed] [Google Scholar]; b) Wang Y., Torres J. A., Stieg A. Z., Jiang S., Yeung M. T., Rubin Y., Chaudhuri S., Duan X., Kaner R. B., ACS Nano 2015, 9, 9486. [DOI] [PubMed] [Google Scholar]
  • 7. Nguyen N. N., Lee H. C., Kang B., Jo M., Cho K., Nano Lett. 2019, 19, 1758. [DOI] [PubMed] [Google Scholar]
  • 8. Kratzer M., Teichert C., Nanotechnology 2016, 27, 292001. [DOI] [PubMed] [Google Scholar]
  • 9.a) Ashraf A., Wu Y., Wang M. C., Yong K., Sun T., Jing Y., Haasch R. T., Aluru N. R., Nam S., Nano Lett. 2016, 16, 4708; [DOI] [PubMed] [Google Scholar]; b) Hong G., Han Y., Schutzius T. M., Wang Y., Pan Y., Hu M., Jie J., Sharma C. S., Muller U., Poulikakos D., Nano Lett. 2016, 16, 4447; [DOI] [PubMed] [Google Scholar]; c) Kim C.‐H., Kymissis I., J. Mater. Chem. C 2017, 5, 4598. [Google Scholar]
  • 10. Kumar A., Banerjee K., Dvorak M., Schulz F., Harju A., Rinke P., Liljeroth P., ACS Nano 2017, 11, 4960. [DOI] [PubMed] [Google Scholar]
  • 11.a) Sirringhaus H., Adv. Mater. 2014, 26, 1319; [DOI] [PMC free article] [PubMed] [Google Scholar]; b) Schuettfort T., Watts B., Thomsen L., Lee M., Sirringhaus H., McNeill C. R., ACS Nano 2012, 6, 1849; [DOI] [PubMed] [Google Scholar]; c) Khim D., Luzio A., Bonacchini G. E., Pace G., Lee M.‐J., Noh Y.‐Y., Caironi M., Adv. Mater. 2018, 30, 1705463; [DOI] [PubMed] [Google Scholar]; d) Soeda J., Okamoto T., Hamaguchi A., Ikeda Y., Sato H., Yamano A., Takeya J., Org. Electron. 2013, 14, 1211; [Google Scholar]; e) He T., Wu Y., D'Avino G., Schmidt E., Stolte M., Cornil J., Beljonne D., Ruden P. P., Würthner F., Frisbie C. D., Nat. Commun. 2018, 9, 2141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.a) Venugopal A., Chan J., Li X., Magnuson C. W., Kirk W. P., Colombo L., Ruoff R. S., Vogel E. M., J. Appl. Phys. 2011, 109, 104511; [Google Scholar]; b) Lee E., Lee H. C., Jo S. B., Lee H., Lee N.‐S., Park C. G., Lee S. K., Kim H. H., Bong H., Cho K., Adv. Funct. Mater. 2016, 26, 562. [Google Scholar]
  • 13. Gobbi M., Pietrobon L., Atxabal A., Bedoya‐Pinto A., Sun X., Golmar F., Llopis R., Casanova F., Hueso L. E., Nat. Commun. 2014, 5, 4161. [DOI] [PubMed] [Google Scholar]
  • 14. Braun S., Salaneck W. R., Fahlman M., Adv. Mater. 2009, 21, 1450. [Google Scholar]
  • 15. Witte G., Lukas S., Bagus P. S., Wöll C., Appl. Phys. Lett. 2005, 87, 263502. [Google Scholar]
  • 16. Gaul C., Hutsch S., Schwarze M., Schellhammer K. S., Bussolotti F., Kera S., Cuniberti G., Leo K., Ortmann F., Nat. Mater. 2018, 17, 439. [DOI] [PubMed] [Google Scholar]
  • 17. Ojeda‐Aristizabal C., Santos E. J. G., Onishi S., Yan A., Rasool H. I., Kahn S., Lv Y., Latzke D. W., Velasco J. Jr., Crommie M. F., Sorensen M., Gotlieb K., Lin C. Y., Watanabe K., Taniguchi T., Lanzara A., Zettl A., ACS Nano 2017, 11, 4686. [DOI] [PubMed] [Google Scholar]
  • 18. Senyavin V. M., Davydov V. A., Kashevarova L. S., Rakhmanina A. V., Agafonov V., Allouchi H., Céolin R., Sagon G., Szwarc H., Chem. Phys. Lett. 1999, 313, 421. [Google Scholar]
  • 19. Sundqvist B., Sci. Rep. 2015, 4, 6171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20. Zou Y., Liu B., Wang L., Liu D., Yu S., Wang P., Wang T., Yao M., Li Q., Zou B., Cui T., Zou G., Wågberg T., Sundqvist B., Mao H.‐K., Proc. Natl. Acad. Sci. USA 2009, 106, 22135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Fagerström J., Stafström S., Phys. Rev. B 1996, 53, 13150. [DOI] [PubMed] [Google Scholar]
  • 22. Zhao W., Zhou W.‐L., Chen L.‐Q., Huang Y.‐Z., Zhang Z.‐B., Fung K. K., Zhao Z.‐X., J. Solid State Chem. 1994, 112, 412. [Google Scholar]
  • 23. Kim K., Lee T. H., Santos E. J. G., Jo P. S., Salleo A., Nishi Y., Bao Z., ACS Nano 2015, 9, 5922. [DOI] [PubMed] [Google Scholar]
  • 24. Mirzayev R., Mustonen K., Monazam M. R. A., Mittelberger A., Pennycook T. J., Mangler C., Susi T., Kotakoski J., Meyer J. C., Sci. Adv. 2017, 3, e1700176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Kaur N., Dharamvir K., Jindal V. K., Chem. Phys. 2008, 344, 176. [Google Scholar]
  • 26.a) Venables J. A., Spiller G. D. T., Hanbucken M., Rep. Prog. Phys. 1984, 47, 399; [Google Scholar]; b) Venables J. A., Surf. Sci. 1994, 299–300, 798. [Google Scholar]
  • 27. Venables J. A., Brune H., Phys. Rev. B 2002, 66, 195404. [Google Scholar]
  • 28. Bommel S., Kleppmann N., Weber C., Spranger H., Schafer P., Novak J., Roth S. V., Schreiber F., Klapp S. H., Kowarik S., Nat. Commun. 2014, 5, 5388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Liu H., Reinke P., J. Chem. Phys. 2006, 124, 164707. [DOI] [PubMed] [Google Scholar]
  • 30. Virkar A., Mannsfeld S., Oh J. H., Toney M. F., Tan Y. H., Liu G. y., Scott J. C., Miller R., Bao Z., Adv. Funct. Mater. 2009, 19, 1962. [Google Scholar]
  • 31.a) Kobayashi S., Takenobu T., Mori S., Fujiwara A., Iwasa Y., Appl. Phys. Lett. 2003, 82, 4581; [Google Scholar]; b) Haddon R. C., Perel A. S., Morris R. C., Palstra T. T. M., Hebard A. F., Fleming R. M., Appl. Phys. Lett. 1995, 67, 121. [Google Scholar]
  • 32.a) Hlaing H., Kim C. H., Carta F., Nam C. Y., Barton R. A., Petrone N., Hone J., Kymissis I., Nano Lett. 2015, 15, 69; [DOI] [PubMed] [Google Scholar]; b) Kang W., Kitamura M., Arakawa Y., Org. Electron. 2013, 14, 644. [Google Scholar]
  • 33. Yogev S., Matsubara R., Nakamura M., Zschieschang U., Klauk H., Rosenwaks Y., Phys. Rev. Lett. 2013, 110, 036803. [DOI] [PubMed] [Google Scholar]
  • 34. Sinha D., Lee J. U., Nano Lett. 2014, 14, 4660. [DOI] [PubMed] [Google Scholar]
  • 35.a) Jnawali G., Rao Y., Beck J. H., Petrone N., Kymissis I., Hone J., Heinz T. F., ACS Nano 2015, 9, 7175; [DOI] [PubMed] [Google Scholar]; b) Qin S., Chen X., Du Q., Nie Z., Wang X., Lu H., Wang X., Liu K., Xu Y., Shi Y., Zhang R., Wang F., ACS Appl. Mater. Interfaces 2018, 10, 38326. [DOI] [PubMed] [Google Scholar]
  • 36.a) Liu Y., Stradins P., Wei S. H., Angew. Chem., Int. Ed. 2016, 55, 965; [DOI] [PubMed] [Google Scholar]; b) Liu Y., Stradins P., Wei S.‐H., Sci. Adv. 2016, 2, e1600069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Li Z. Q., Henriksen E. A., Jiang Z., Hao Z., Martin M. C., Kim P., Stormer H. L., Basov D. N., Nat. Phys. 2008, 4, 532. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information


Articles from Advanced Science are provided here courtesy of Wiley

RESOURCES