Home Carbon Nanomembranes
Article Publicly Available

Carbon Nanomembranes

  • Polina Angelova EMAIL logo and Armin Gölzhäuser
Published/Copyright: March 18, 2017
Become an author with De Gruyter Brill

Abstract

This chapter describes the formation and properties of one nanometer thick carbon nanomembranes (CNMs), made by electron induced cross-linking of aromatic self-assembled monolayers (SAMs). The cross-linked SAMs are robust enough to be released from the surface and placed on solid support or over holes as free-standing membranes. Annealing at ~1000K transforms CNMs into graphene accompanied by a change of mechanical stiffness and electrical resistance. The developed fabrication approach is scalable and provides molecular level control over thickness and homogeneity of the produced CNMs. The mechanisms of electron-induced cross-linking process are discussed in details. A variety of polyaromatic thiols: oligophenyls as well as small and extended condensed polycyclic hydrocarbons have been successfully employed, demonstrating that the structural and functional properties of the resulting nanomembranes are strongly determined by the structure of molecular monolayers. The mechanical properties of CNMs (Young’s modulus, tensile strength and prestress) are characterized by bulge testing. The interpretation of the bulge test data relates the Young’s modulus to the properties of single molecules and to the structure of the pristine SAMs. The gas transport through the CNM is measured onto polydimethylsiloxane (PDMS) - thin film composite membrane. The established relationship of permeance and molecular size determines the molecular sieving mechanism of permeation through this ultrathin sheet.

Carbon nanomembranes (CNMs) [1] are very thin (~1 nm), synthetic two-dimensional (2D) layers or sheets with tailored physical, chemical or biological function. With their two opposing surfaces they interface and link different environments by their distinct physical and chemical properties, which depend on the thickness, molecular composition, structure and environment on either side. Due to their minute nanometer thickness and 2D architecture, they can be regarded as “surfaces without bulk” separating regions with different gaseous, liquid or solid components and controlling any exchange of materials between them [2].

CNMs have an extremely large surface-to-volume ratio but at the same time are robust enough to stand freely in air and to be prepared as freestanding structures (cf. Figure 1). Their uniqueness comes from the combination of their molecular-scale thickness and their macroscopic dimensions, thus linking the properties of the macroscopic materials with those of individual molecules. CNMs are predicted to possess a superior performance [3] and withstand pressure of a few bar without rupturing. This allows their use in technical applications, such as force sensing, and separation of materials where CNMs with pores of controlled size allow a faster passage of the selected gas or liquid molecules than any conventional filter [4].

Figure 1: Helium ion microscopy image of a CNM, placed over a copper transmission electron microscopy hexagonal mesh grid: (a) field of view 70 μm, (b) field of view 2.25 mm.
Figure 1:

Helium ion microscopy image of a CNM, placed over a copper transmission electron microscopy hexagonal mesh grid: (a) field of view 70 μm, (b) field of view 2.25 mm.

Although very young, the group of functional supported or freestanding 2D nanolayers is already large. Among those, graphene (Figure 2(c)) [5] that only consists of a single layer of carbon (only 3 Å) is by far the thinnest and structurally most simple one. On the other side, biomembranes (7–9 nm), consisting of lipid bilayers with embedded membrane proteins, are probably the most complex nanolayers that perform numerous tasks of living cells. Figure 2 shows a scheme with these “extreme” complex and simple nanolayers placed on the left and right, respectively. Other nanomembranes – whether of artificial, i.e., synthetic or engineered, or of biological origin – lay between these cases.

Figure 2: Schematic presentation of ultrathin carbon-rich nanomembranes: (a) biological membrane, (b) the route from self-assembled monolayer through CNM to graphene [6], (c) atomically thick graphene sheet.
Figure 2:

Schematic presentation of ultrathin carbon-rich nanomembranes: (a) biological membrane, (b) the route from self-assembled monolayer through CNM to graphene [6], (c) atomically thick graphene sheet.

The structure and function of biological membranes are not in the focus of this chapter. However, one has to keep in mind that biological membranes are, by far, the most functional 2D “machines” on earth. Their applications range from encapsulation and mechanical protection of the cell content to molecular recognition, filtration and purification. Biomembranes can be considered as the “reference system” for functional 2D materials. They demonstrate the manifold of possible membrane application and they benchmark membrane performance.

In the following, we will focus on synthetic, man-made nanomembranes. These can be divided into inorganic, organic and hybrid ones. The inorganic membranes possess superior chemical and thermal stabilities and offer higher fluxes in contrast to polymer materials. One of the central positions among the inorganic membranes belongs to semiconducting silicon-based nanomembranes [7]. They combine thickness of 10–50 nm with extreme flexibility and robustness, thus allowing their incorporation in stretchable electronic devices. An intermediate class, which received deserved attention in the last decade, is the organic–inorganic hybrid membranes that include relatively new classes of porous materials like the organic–inorganic interpenetrated networks (IPNs) [8], metal–organic frameworks [9] and mixed matrix membranes [10]. They carry the macroscopic robustness of the ceramic materials and the flexibility, light weight and economical processability of polymers. Technologically important class of nanomembranes is the polymeric membranes [11]. Scaled up to industrial needs, they are mostly utilized in water and gas separation processes and biofiltration in medicine. But despite their easy preparation, low price and propensity to versatile chemical functionalizations, they lack chemical and mechanical stability. They also suffer from aging and fouling by time and to date their thickness cannot be reduced below 8 nm [12], thus limiting their permeance performance. There are two main processes of fabrication of ultrathin organic nanomembranes. Since the 1990s the layer-by-layer (LbL) technique is employed to form polymer membranes for corrosion protection, sensing and drug delivery [11a, 13]. In the LbL process, an electrically charged surface is sequentially dipped into positively and negatively charged polyelectrolytes, leading to the formation of polymeric membranes of well-defined molecular composition with thicknesses from ~15 nm to several hundreds of nanometers. An essential characteristic of this method is that it offers easy control over the thickness. Thin polymer membranes are also obtained by interfacial polymerization [12, 14], including air/liquid interface and liquid/liquid interface.

In the literature the term “nanomembrane” often denotes sheets with thicknesses below 100 nm. However, in this chapter we treat much thinner nanomembranes, namely those with a thickness of one molecule. The synthesis of freestanding nanomaterials with an atomic or molecular thickness is a challenging task and to date only a scarce number of chemical pathways have been developed. Atomically thin membranes can be made by exfoliating single sheets out of a layered material [15]. Graphene, consisting of a few (~1–5) layers of carbon atoms, was initially made by exfoliation [5] and is nowadays produced by a variety of techniques [16], allowing researchers to gain new insights into physics and chemistry in two dimensions. However, the surface of graphene is homogeneous and chemically inert, which is unfavorable for an efficient surface functionalization. On the contrary, covalent organic frameworks (COFs) [17] of arylalkynes [18] and related 2D systems [19] are nanomembranes with heterogeneous, chemically reactive surfaces. The first step in their creation is the synthesis of organic molecules of well-defined size and shape, with functionalities at defined positions. When brought into proximity, adjacent molecules form multiple covalent bonds and assemble in a 2D lattice. The second stage of nanomembrane fabrication is detaching the carbon nanolayer from its initial or sacrificial surface. A common feature of all these synthetic approaches is that they all are driven by spontaneous self-assembly. This quite simple and energy-saving approach, inspired by the biological membranes and mechanisms in living organisms, has gained wide applicability due to its superiority in tailoring organization with molecular-level precision. However, the necessity of specially designed molecular precursors requires high skills in organic synthesis and is time-consuming. Most of the described approaches are also restricted to micrometer lateral dimensions, expensive and sophisticated fabrication and low mechanical stability.

A molecular route toward the large-scale fabrication of a CNM that combines the thinness of graphene with the chemical functionality of a COF and the ease of fabrication of LbL films was described by Angelova et al. [6]. The fabrication route (Figure 3(a)) utilizes a sequence of (i) molecular assembly on a solid surface, (ii) radiation-induced 2D polymerization and (iii) a liftoff of the network of cross-linked molecules [20]. Figure 3(b) and (c) shows freestanding CNMs, formed from hexa-peri-benzocoronene (HBC) derivatives (Figure 9, precursor 3b and 3c), which, after etching of the initial gold film, are transferred [20] onto a perforated support (copper TEM grid) and imaged with a helium ion microscope (HIM) [21]. This novel charged particle microscopy combines high resolution (~4 Å) with high surface sensitivity and the possibility to image nonconducting ultrathin specimens [22].

Figure 3: (a) Schematics of the structure and fabrication of a CNM, (b and c) HIM images of CNM, prepared from polyaromatic (HBC, precursors 3b and 3c, presented in Figure 9) molecules.
Figure 3:

(a) Schematics of the structure and fabrication of a CNM, (b and c) HIM images of CNM, prepared from polyaromatic (HBC, precursors 3b and 3c, presented in Figure 9) molecules.

Thereby the surface-bound self-assembled monolayers (SAMs) are converted into freestanding CNMs, which are structurally amorphous, chemically durable [23] and thermally [24] and mechanically robust [20] to be suspended over orifices of micrometer sizes. Furthermore, upon annealing they can be converted into well-conducting nanocrystalline graphene sheets (Figure 4) [25]. In addition, the freestanding CNM has two faces, the chemistry of which is determined by the surface-active and the end group of the molecular precursors. Thus a smart molecular design can easily tailor the chemical functionalities on both faces of the membrane, and by means of a 2D synthesis a refashioning of its properties and functions is easy to achieve [26]. But the most remarkable aspect of this pathway is that the thickness, porosity and surface chemistry of the nanomembranes are determined by the molecular order of the monolayers, and structural and functional features are passed on from the molecules through their monolayers to the CNMs and finally on to the graphene.

Figure 4: High-resolution TEM images of a CNM (a) before and (b) after annealing in vacuum at 1,200 K. Reprinted with permission from Ref. [25]. Copyright (2011) American Chemical Society.
Figure 4:

High-resolution TEM images of a CNM (a) before and (b) after annealing in vacuum at 1,200 K. Reprinted with permission from Ref. [25]. Copyright (2011) American Chemical Society.

1 Molecular Mechanisms of Electron-Induced Cross-Linking

It has been shown that electron, ion and photon beams induce specific chemical reactions, i.e., bond dissociation, oxidation/reduction or polymerization, in SAMs that depend strongly on their building blocks – the molecules. The response of a SAM to low-energy electron irradiation depends on the chemical structure of the ordered molecules. In general the aliphatic monolayers are heavily destroyed and thus broadly utilized as positive tone resists [27], but the phenyl-based SAMs are stabilized through the formation of intermolecular cross-links and can be used as negative resists. In aliphatic SAMs a cleavage of C–H bonds through resonant and nonresonant processes is induced, which leads to an orientational and conformational disorder of the chains, desorption of material and formation of C=C double bonds in the fragments, remaining on the surface, whereas in aromatic SAMs, it was found that upon irradiation with electrons, C–H cleavage occurs, which is then followed by cross-linking between neighboring phenyl units. During this process, cross-linked molecules maintain their preferred orientation and almost no material desorbs. Hence the electrons generate a molecularly dense monolayer that could be used as a negative tone electron resist [23, 27, 28].

In addition, electron irradiation also changes the surface chemistry of the nanomembrane. Several groups have investigated the chemical transformation of nitro-, cyano-, trifluoromethyl- para-substituents upon electron irradiation [29, 30]. While the underlying aromatic rings are partially dehydrogenated, the hydrogen atoms liberated from the aromatic cores locally reduce the nitro and cyano groups to amino and aminomethyl groups, which can be further chemically modified by electrophilic agents [26a, 26b, 31]. By employing a local electron exposure, one can define surface regions that can be used as a template for a site-selective molecular immobilization. This process is named chemical lithography, as the lithographic exposure directly affects the surface chemistry of the SAM [29a, 32].

Resolving the molecular mechanisms of the e-beam-induced polymerization in aromatic SAMs is a complicated task, since the long-range order in the cross-linked layers is distorted and a broad spectrum of molecular species can be formed. Turchanin et al. [33] employed complementary spectroscopic techniques such as X-ray and UV photoelectron spectroscopy (XPS and UPS), near-edge X-ray absorption fine structure spectroscopy (NEXAFS) and thermal desorption spectroscopy (TDS) to investigate the mechanisms of cross-linking in biphenyl-4-thiol (BPT) SAMs on gold (Figure 5). The experimental data are completed by quantum chemical calculations that allow the derivation of a model for the e-beam-induced chemical and structural transformations in aromatic thiol SAMs and agree well with the response to electron irradiation of other oligophenylene-based SAMs [29b, 29c, 34].

Figure 5: Schematic presentation of cross-linking of biphenylthiol-based monolayers: (a) low-energy electron irradiation and emission of secondary electrons, (b) dissociation of C–H bonds, (c) cross-linked molecules.
Figure 5:

Schematic presentation of cross-linking of biphenylthiol-based monolayers: (a) low-energy electron irradiation and emission of secondary electrons, (b) dissociation of C–H bonds, (c) cross-linked molecules.

When a molecular layer is exposed to low-energy electrons, the electrons are effectively captured by the ordered molecular film, followed by a partial decomposition, desorption of hydrogen and molecular fragments, orientational and conformational disordering, and intermolecular cross-linking. For alkanethiol SAMs on gold [35], it was found that the electron energies, resonant with the electron orbitals of the aliphatic components, have a maximum at ~10 eV with a cross section of 10–16 cm2. Therefore, dissociative electron attachment (DEA) is most likely the dominating process. DEA [36] is observed when a free electron interacts with an atom or molecule and is temporally captured to form a negative ion resonance (also called transition negative ion). This electronically excited anionic state, in which the attached electron resides on one of the unoccupied molecular orbitals, decays in a release of hydrogen anion and formation of carbon radical centers within the film or a release of hydrogen atom and formation of molecular anionic species. These reactive centers further react with the adjacent phenyl rings to form intermolecular C–C covalent bonds as the delocalization of π-electrons over the δ-framework of the aromatic ring retains its integrity through the irradiation process. Desorption of hydrogen molecules is generated through scattering of the formed hydride/hydrogen atom within the dense molecular layer.

For BPT SAMs ~650 primary electrons per molecule with energy of 50 eV are necessary to create the cross-linked molecular network [33]. It has been shown that secondary electrons (SEs) generated in the substrate [37] also contribute to the cross-linking along with the primary e-beam (Figure 5) [38]. Taking into account the SE yield [39], the cross section for the cross-linking of BPT was estimated to be ~10–18 cm2. This number is an average value without specifying an exact number of the involved chemical bonds and their chemical nature.

Electron-induced processes in polycyclic aromatic hydrocarbons were studied in the gas phase [40]. Resonances, characteristic for DEA, were observed at energies of 7–8 eV with effective cross sections of 10–17–10–16 cm2. As these cross sections are close in magnitude to the one found for biphenyl cross-linking, it is likely that DEA contributes to the C–H cleavage in BPT SAMs.

In Figure 6, XPS, UPS and NEXAFS data for a BPT SAM before and after irradiation with 50 eV electrons and a dose of 45 mC/cm2 are presented. A comparison of these data reveals specific transformations upon cross-linking of the carbonaceous part of the monolayer and of the sulfur/gold interface. While the XP C1s signal (cf. Figure 6(a)) mostly preserves its initial intensity, demonstrating small reduction in the amount of carbon atoms after electron irradiation, the corresponding data of the C1s X-ray absorption show a reduction of the π* resonance intensity while at the same time the δ* resonances become more intense (cf. Figure 6(d)). The width of all resonances increases; however, their positions remain unchanged. These observations are attributed to a partial loss of aromaticity within the monolayer due to the interconnection of adjacent molecules via C–C bonds or even due to their partial decomposition. Measurements of angular dependencies of the resonance intensities of the pristine and e-beam-irradiated SAMs revealed an increase of the averaged molecular tilt within the monolayer upon cross-linking from ϕ = 31° to 41°.

Figure 6: Spectroscopic characterization of the e-beam-induced cross-linking of a biphenyl-4-thiol SAM (e-beam energy of 50 eV, electron dose of 45 mC/cm2): (a) X-ray photoelectron spectroscopy (XPS), (monochromatic Al Ka, 1,486.7 eV), (b) temperature desorption spectroscopy (TDS), (c) ultraviolet photoelectron spectroscopy (UPS), (He I, 21.22 eV), (d) near-edge X-ray absorption fine structure spectroscopy (NEXAFS) (schematic representation of a pristine BPT SAM), (e) an e-beam cross-linked BPT SAM. Reprinted from Ref. [1]. Copyright (2012) with permission from Elsevier.
Figure 6:

Spectroscopic characterization of the e-beam-induced cross-linking of a biphenyl-4-thiol SAM (e-beam energy of 50 eV, electron dose of 45 mC/cm2): (a) X-ray photoelectron spectroscopy (XPS), (monochromatic Al Ka, 1,486.7 eV), (b) temperature desorption spectroscopy (TDS), (c) ultraviolet photoelectron spectroscopy (UPS), (He I, 21.22 eV), (d) near-edge X-ray absorption fine structure spectroscopy (NEXAFS) (schematic representation of a pristine BPT SAM), (e) an e-beam cross-linked BPT SAM. Reprinted from Ref. [1]. Copyright (2012) with permission from Elsevier.

The UPS data show photoemission bands at 2.7 and 4.4 eV as well as a loss of fine structures around 7–10 eV upon cross-linking (Figure 6(c)). Based on quantum chemical calculations of the electronic structure, which were carried out for numerous possible cross-linked BPT dimers, the features in the UPS can be attributed to electronic excitations and thus allow the identification of individual molecular species such as BBDS, B1BPT, B4BPT, BBPDT and BBPDTn (Table 1). For the analysis of the photoemission lines at first the molecular energy levels were calculated and then the allowed dipole transitions were considered [41]. The comparison of calculated and measured UPS suggests a prevailing formation of BBPDT and BBPDTn species upon e-beam irradiation.

Table 1:

Molecular species whose formation upon cross-linking of a BPT SAM was considered and for which UPS and HOMO–LUMO gaps were calculated.

Table 1: Molecular species whose formation upon cross-linking of a BPT SAM was considered and for which UPS and HOMO–LUMO gaps were calculated.

This conclusion is consistent with the thermal desorption of biphenylene fragments (m/z = 151) in a partially cross-linked BPT SAM, which was observed by TDS (Figure 6(b)). The appearance of dimers is also consistent with a time of flight secondary ion mass spectrometry study of irradiated 4ʹ-methyl-1,1ʹ-biphenyl-4-thiol (MBPT) SAMs on gold by Cyganik et al. [34a].

Distinct changes upon e-beam-induced cross-linking take place at the sulfur/gold interface of a BPT SAM. Figure 6(a) shows that besides thiolate SAM species [42] with a S2p3/2 binding energy (BE) of 162.0 eV, new sulfur species with a BE of 163.5 eV form in a cross-linked SAM. Although its BE coincides with that of thioethers (R–S–R) or organodisulfides (R–S–S–R) [43], the presence of other species cannot be unequivocally excluded. It has been demonstrated that metal surfaces are not that rigid and actually reveal a substantial rearrangement upon formation of thiol-based SAMs [44]. In addition to vacancy depressions or ad-islands, which were attributed to a stress release [45], formation of thiolate-dimers (RS–Au–SR) also takes place [46, 47] and the measured S2p doublet at 162.0 eV is attributed to these sulfur species. It was shown that some molecules within thiol-SAMs form pairs, stabilized by an additional gold atom between neighboring sulfur atoms as depicted schematically in Figure 7(a). The density functional theory (DFT) - based geometry optimization (carried out for the related bisbiphenyl-disulfanyl species (cf. Table 1)) [33] revealed a substantial outward tilting of the aromatic backbones (see middle of Figure 7(a)) due to steric repulsion of the lower phenyl rings, which is in a qualitative agreement with the observed by NEXAFS increase in the tilt of the aromatic backbone upon cross-linking from ϕ = 31° to 41° [33]. Since cross-linked BPT molecules with multiple cross-links are less separated, they may provide sufficient space for formation of thiolate-Au-thiolate pairs with a higher downward tilt. These pairs can be stabilized by additional links at the upper phenyl rings as shown in Figure 7(a), middle.

Figure 7: Schematic summary of the structural properties of pristine and cross-linked BPT SAMs: (a) molecular species formed upon e-beam irradiation, (b) conversion of a pristine BPT SAM into cross-linked BPT SAM. Reprinted with permission from Ref. [1]. Copyright (2012) with permission from Elsevier.
Figure 7:

Schematic summary of the structural properties of pristine and cross-linked BPT SAMs: (a) molecular species formed upon e-beam irradiation, (b) conversion of a pristine BPT SAM into cross-linked BPT SAM. Reprinted with permission from Ref. [1]. Copyright (2012) with permission from Elsevier.

To realize a cross-linking in two dimensions and the formation of CNMs, it is important to consider the molecular packing motifs adopted in the pristine BPT SAMs (Figure 7(b), left) [48, 49]. Isolated BPT molecules reveal a characteristic torsional angle between the upper and lower phenyl rings. Within the crystalline SAM both phenyl rings may not be coplanar but still have a substantial twisting. Theoretical studies indicate that this molecular degree of freedom is of key importance for the formation of ordered films of oligophenylene-based SAMs [50]. Moreover, in view of the herringbone arrangement, which is adopted in the crystalline phase of oligophenylenes [51], a lateral cross-linking occurring upon BBPDT (Table 1) formation comprises a molecular rearrangement, such as a rotation. Therefore, the formation of extended chains exhibiting the double Ph–Ph bonding motif (Figure 7(a), left) is sterically unfavorable and a 2D network can propagate via molecules that are linked either at the upper or lower ring as schematically shown in Figure 7(a) middle.

A further consequence of the cross-linking is a reduction of the spatial separation of neighboring BPT molecules. While the intermolecular distances within the pristine film are essentially given by the van der Waals dimensions of the molecules, the irradiation-induced additional carbon–carbon links also enable shorter distances. The density of the cross-linked films increases locally and may lead to a formation of “nano-voids,” containing isolated and noncross-linked molecules (denoted as dark molecules in the scheme in Figure 7(b), right).

Due to the multiple cross-links the CNM exhibits a thermal stability up to 1,000 K [24]. Pristine biphenylthiol SAMs desorb above 400 K (~130°C) (cf. Figure 6(b)), whereas e-beam-irradiated SAMs become more stable with an increasing irradiation dose, i.e., with an increasing degree of cross-linking. This process saturates at a dose of ~50 mC/cm2 for 50 eV electrons. For such samples only a slight (~10 %) reduction of their carbon XP signal is observed even upon annealing at ~1,000 K. But the data show that the sulfur atoms initially present in the monolayer, and in the CNM, continuously desorb upon heating through cleavage of C–S bonds until they completely vanish at temperatures above 800 K (cf. Figure 6(b)). The absence of any sulfur signal in the cross-linked monolayer after annealing at elevated temperatures demonstrates that the remaining carbonaceous film is not anchored by thiolate but is solely stabilized by covalent bonds within the aromatic network, which is directly coupled to the gold surface via van der Waals interactions. This conclusion is well corroborated by the temperature-dependent NEXAFS measurements [33], revealing a downward tilting of the aromatic rings and a substantial broadening of the π* resonances at elevated temperatures. As demonstrated by quantum chemical calculations, the saturation of the cross-linking process at a dose of ~50 mC/cm2 can be attributed to the electronic quenching of the anionic molecular states by means of tunneling or hopping (Figure 7).

Upon electron irradiation of an aromatic monolayer, a dense molecularly thick network is formed, which is robust enough to be further removed from its initial substrate (Au [20], Si3N4 [30], Cu [52]) and placed onto solid or perforated support (Figure 8) as the thickness of the resulting CNM is adopted by the thickness of its parenting SAMs. Motivated by these findings, we applied similar protocols to a variety of other polyaromatic molecules, aiming to examine the change in structural features and functionality of the membrane. Figure 9 shows schematic drawing of the molecular route, applied to different types of monolayers.

Figure 8: (a) Interference contrast of transferred CNM onto Si substrate with 300 nm silicon oxide layer, (b) helium ion microscopy image of a freestanding CNM (in blue color), transferred onto copper TEM hexagonal mesh grid.
Figure 8:

(a) Interference contrast of transferred CNM onto Si substrate with 300 nm silicon oxide layer, (b) helium ion microscopy image of a freestanding CNM (in blue color), transferred onto copper TEM hexagonal mesh grid.

Figure 9: Schematic for the formation of CNMs and graphene from various molecular precursors: (a–c) schematic illustration of the fabrication route for CNMs and graphene. SAMs are prepared on a substrate (i), then cross-linked by electron irradiation to form CNM of a monomolecular thickness (ii). The CNM can be released from the underlying substrate (iii) and transferred onto a new substrate as a freestanding or supported material. Annealing to 900°C transforms CNM into graphene. (a) Fabrication of atomically thin CNMs and graphene from precursors 1a–c in (d), (b) fabrication of thicker CNMs and few layer graphene sheets from precursors 2a–f in (d), (c) fabrication of CNMs and graphene sheets with nanopores from precursors 3a–c in (d), (d) chemical structures of the molecular precursors used for preparation of CNMs and graphene. Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.
Figure 9:

Schematic for the formation of CNMs and graphene from various molecular precursors: (a–c) schematic illustration of the fabrication route for CNMs and graphene. SAMs are prepared on a substrate (i), then cross-linked by electron irradiation to form CNM of a monomolecular thickness (ii). The CNM can be released from the underlying substrate (iii) and transferred onto a new substrate as a freestanding or supported material. Annealing to 900°C transforms CNM into graphene. (a) Fabrication of atomically thin CNMs and graphene from precursors 1a–c in (d), (b) fabrication of thicker CNMs and few layer graphene sheets from precursors 2a–f in (d), (c) fabrication of CNMs and graphene sheets with nanopores from precursors 3a–c in (d), (d) chemical structures of the molecular precursors used for preparation of CNMs and graphene. Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.

2 Tuning of CNM’s Properties on a Molecular Level

Various types of thiol-based precursors were studied on Au (111) polycrystalline substrates: nonfused oligophenyl derivatives (Figure 9(d), structures 1a–c), which possess linear molecular backbones providing an improved structural ordering of the formed SAMs; condensed polycyclic precursors like naphthalene (NPTH), anthracene (ANTH) and pyrene (MP) derivatives (Figure 9(d), structures 2a–f), which are more rigid and should result in a higher stability and an increased carbon density of the monolayers; “bulky” molecules, like the noncondensed hexaphenylbenzene derivative with a propeller-like structure (structure 3a) and extended disk-type polycyclic aromatic hydrocarbons such as HBC derivatives (Figure 9(d), structures 3b–c) [53]. The former are equipped with long alkyl chains and a surface-active group, which is attached to the π-conjugated backbone through a flexible methylene linker. This molecular structure enables a control over the thickness and packing density of the SAMs by varying the conditions of preparation. In Figure 9(a) oligophenyls with a linear molecular backbone form well-ordered monolayers that can be cross-linked into homogeneous CNMs. After pyrolysis, the CNMs transform into graphene, whose thickness depends on the density of carbon atoms in the monolayers. In Figure 9(b), condensed polycyclic precursors also form monolayers that are cross-linked into CNMs. After pyrolysis, the CNMs transform into graphene, whose thickness is higher than that in structure 1a, Figure 9(d), even if the carbon density is the same as in Figure 9(a). Figure 9(c) shows bulky aromatic hydrocarbons that assemble in a less-ordered monolayer and polymerize into CNMs with pores. After annealing, these nanomembranes transform into thicker graphene sheets with pores. Hence, the produced graphene adopts features from the preceding CNM, which itself adopts features from the monolayer, i.e., from molecules and surface.

SAMs form due to formation of strong bonds between the sulfur and the gold atoms, which is accompanied by van der Waals interactions between the carbon atoms. To obtain SAMs with a desired molecular packing, one can adjust parameters like immersion time, temperature, concentration and polarity of the solvents. XPS data of aromatic SAMs, representing diverse types of precursors, are shown in Figure 10 (left). For each SAM, its chemical composition and thickness can be derived from the BE and intensity of the C1s, S2p and Au4f photoelectron signals. The sulfur signal consists of a doublet with a S2p3/2 BE of 162.0 eV, which unambiguously demonstrates the formation of sulfur–gold bonds [54].

Figure 10: XPS data of pristine SAMs and CNMs: (a–f) XP spectra of C1s and S2p signals of the pristine (in left) and electron-irradiated monolayers (in right). In the insets are presented the chemical structures of the monolayer precursors. Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.
Figure 10:

XPS data of pristine SAMs and CNMs: (a–f) XP spectra of C1s and S2p signals of the pristine (in left) and electron-irradiated monolayers (in right). In the insets are presented the chemical structures of the monolayer precursors. Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.

Only for the HBC derivatives (structures 3b–c), the presence of a second sulfur species with the BE of the S2p3/2 signal at 163.6 eV is observed. This signal originates from physisorbed molecules and/or of disulfides [54], which may be stabilized by π–π interactions between the large aromatic cores. Aromatic and aliphatic carbons contribute to the C1s signal at BEs of ~284.2 and ~285.0 eV, respectively [55]. Stoichiometry and thickness as obtained from XPS correspond to the composition of the precursor molecules and indicate the formation of SAMs with an “upright” molecular orientation. By varying the precursors, the thickness of the aromatic monolayers can be adjusted from ~6 Å for NPHT (Figure 9(d): structure 2a) to ~24 Å for HBC-CN (Figure 9(d): structure 3c), which directly correlates with their molecular lengths (Figure 10 and Table 2).

Table 2:

Effective thickness of pristine SAMs and CNMs; carbon reduction upon electron irradiation.

SampleThickness SAM [Å]Thickness CNM [Å]Reduction of C [%]
1a (BPT)1095
1b (BP3)121016
1c (TPT)13124
2a (NPHT)669
2b (ATRH)992
2c (1MP)9810
2d (MP1)984
2g (MP3)11108
2f (MP5)10811
3a (HPB)8813
3b1 (HBC-Br)10104
3b2 (HBC-Br)12115
3b3 (HBC-Br)19172
3c1 (HBC-CN)12106
3c2 (HBC-CN)14125
3c3 (HBC-CN)24223

*Different conditions were applied for preparation of SAMs from 3b1–3 and 3c1–3.

Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.

In addition, the temperature- and solvent-dependent intermolecular interactions of the HBC derivatives [56] allow one to tune the final SAM thickness by varying the preparation conditions (Table 2).

Scanning tunneling microscopy (STM) and low-energy electron diffraction (LEED) data of three molecular precursors, containing different numbers of carbon atoms per molecule – 2-anthracene (C14), 3-biphenylpropane (C15) and terphenyl (C18) thiols – are shown in Figure 11. The structure and surface density of the SAMs were investigated by STM and LEED. They showed that precursor molecules 1b, 1c, 2a, 2b, 2c and 2e (Figure 9(d)) form well-ordered SAMs with the densely packed (√3 × √3) [57] unit cell of the adsorption places and with the (2√3 × √3) superstructure of the molecular backbones. These structures correspond to a surface area of 26 Å2 per molecule. Note that for these SAMs the surface density of carbon atoms in the monolayer can be precisely tuned by the carbon content of the respective molecular precursors. Short biphenylthiols (Figure 9(d), 1a) exhibit a (2 × 2) arrangement of the adsorption places, which corresponds to a less densely packed monolayer of 28.7 Å2 per molecule [49a]. The formation of LEED patterns and well-ordered SAMs by STM for precursor molecules 2d and 2f and 3a–c (Figure 9(d)) was not verified. As XPS indicates a formation of sulfur–gold bonds for all precursors, we conclude that monolayers of the “bulky” and polycyclic molecules (structures 2d, 2f, 3a–c from Figure 9(d)) are less ordered and probably less densely packed than the monolayers of oligophenyls (1a–c) and the small fused-ring systems (2a–c, 2e).

Figure 11: Structure of pristine SAMs. STM micrographs and LEED patterns (insets) of SAMs from molecular precursor: (a) ANTH, 2b; LEED pattern at 116 eV, (b) BP3, 1b; LEED pattern at 127 eV, (c) TPT, 1c; LEED pattern at 129 eV. For molecular structures see Figure 9(d). Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.
Figure 11:

Structure of pristine SAMs. STM micrographs and LEED patterns (insets) of SAMs from molecular precursor: (a) ANTH, 2b; LEED pattern at 116 eV, (b) BP3, 1b; LEED pattern at 127 eV, (c) TPT, 1c; LEED pattern at 129 eV. For molecular structures see Figure 9(d). Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.

The irradiation of these SAMs with low-energy electrons (50 or 100 eV), using typical doses of ~60 mC/cm2, corresponding to ~3,500 electrons per 1 nm2, leads to a loss of order, as observed in LEED and STM. Figure 10 shows XPS data before and after electron irradiation in the left and right parts, respectively. As seen from the intensities of the C1s and Au4f (not shown), for purely aromatic SAMs the irradiation reduces the carbon content and the monolayer thickness by 5–10 %; in the SAMs that also contain aliphatic chains (structures 1b and 2f from Figure 9(d)) the carbon loss is up to ~16 %.

The obtained freestanding (self-supported) CNMs from the above-mentioned molecular precursors can be seen in Figure 12. It shows HIM micrographs of CNMs from three different types of aromatic molecular precursors (Figure 9(d): molecular structures 1–3). Simply the fact that one can take these images demonstrates that the SAMs of all these molecules have been cross-linked into mechanically stable CNMs. Figure 12(a–c) shows the freestanding CNMs from precursors 1c, 2c and 2d. These HIM micrographs were acquired at different magnifications, demonstrating the successful fabrication of CNMs of various sizes. The field of view in Figure 12(a) is 15 × 15 μm2, which allows to observe some folds in the freestanding 1.2-nm-thick CNM. In the lower left corner of Figure 12(b), the boundary between the freestanding or supported CNM and substrate can clearly be seen. Figure 12(c) shows the field of view of 300 × 300 μm2 with a large and homogeneous CNM of a thickness of 0.8 nm. Macroscopic defects in these nanomembranes are practically negligible on the length scale of these images.

Figure 12: HIM micrographs of freestanding CNMs. After cross-linking the nanomembranes were transferred onto TEM grids. CNMs prepared from (a) TPT, structure 1c in Figure 9(d), suspended over gold TEM grid, (b) MP1, 2d, (c) 1MP, 2c, (d) BP3, 2b, (e) HBC-CN, 3c, (f), HBC-CN, 3c, (g) TPT, 1c, (h) HPB, 3a, in Figure 9(d) the inset shows the histogram of the pore size distribution over the displayed area, (i) freestanding nanocrystalline graphene after pyrolysis of NPHT (2a) CNM, (b–f) CNMs suspended over copper TEM grids, (g–i) CNMs and graphene are suspended over a grid with thin carbon film on Cu grid. Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.
Figure 12:

HIM micrographs of freestanding CNMs. After cross-linking the nanomembranes were transferred onto TEM grids. CNMs prepared from (a) TPT, structure 1c in Figure 9(d), suspended over gold TEM grid, (b) MP1, 2d, (c) 1MP, 2c, (d) BP3, 2b, (e) HBC-CN, 3c, (f), HBC-CN, 3c, (g) TPT, 1c, (h) HPB, 3a, in Figure 9(d) the inset shows the histogram of the pore size distribution over the displayed area, (i) freestanding nanocrystalline graphene after pyrolysis of NPHT (2a) CNM, (b–f) CNMs suspended over copper TEM grids, (g–i) CNMs and graphene are suspended over a grid with thin carbon film on Cu grid. Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.

Since the thickness of CNMs is determined by the precursor molecules and their packing density in SAMs, it can be controlled by tailoring these parameters. Figure 12(e) and (f) displays an example of the thickest CNM (2.2 nm) from precursor 3c, Figure 9(d) at two scales that has about four times higher thickness than the thinnest nanomembrane, prepared from precursor 2a, Figure 9(d) with a thickness of only 0.6 nm (Table 2). An image of an annealed NPHT CNM, transformed into nanocrystalline graphene sheet, can be seen in Figure 12(i). The opportunity to flexibly tune the thickness of CNMs opens broad avenues for the engineering of nanomembranes. A thorough investigation of the surface and structural features of different CNMs by HIM revealed the relation between properties of the precursor molecule, its SAMs and the appearance of the ensuing CNM. If the molecule forms a densely packed SAM (1a–c, 2a–c, 2e in Figure 9(d)), the following CNM is continuous and free of holes. Figure 12(g) shows a high magnification HIM image of a homogeneous CNM made from terphenylthiol. Conversely, CNMs made from HBC (3b–c in Figure 9(d)) or HPB (3a in Figure 9(d)) precursors, two molecules that possess larger sizes and form less well-ordered SAMs, exhibit pores (cf. the HIM images in Figures 12(h) and 13(a–f)). The dark spots in these images are pores that have a very small diameter and a narrow size distribution, as shown in the respective histograms (shown in the insets).

In case of the HBC precursor the mean size of the nanopores is ~6 nm with the surface density of 9.1 × 1014 pores/m2; the more compact HPB precursor shows a size of ~2.4 nm with a surface density of 1.3 × 1015 pores/m2. The formation of nanopores in these CNMs is thus attributed to the large van der Waals radii of HBC and HPB structures and in the case of HBCs to the propensity of the disk-like molecules for intermolecular stacking, which competes with the molecule–substrate interactions and lowers the homogeneous coverage in the respective SAMs. We also observe that the average pore diameter of the HBC-based CNMs decreases from 6.4 to 2.5 nm when the SAM thickness increases from 1 to 2 nm (Figure 13).

Figure 13: HIM images of freestanding CNMs: (a–c) prepared from HBC-Br (Figure 9(d), 3b) monolayers of different thickness, (d–f) prepared from HBC-CN (3c) monolayers of different thickness. The thickness of the membranes is written at the top right corner of each image. The histograms in the insets show the distribution of the pore diameter with a step of 0.5 nm. The membranes are placed onto Quantifoil TEM grids type Multi A. Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.
Figure 13:

HIM images of freestanding CNMs: (a–c) prepared from HBC-Br (Figure 9(d), 3b) monolayers of different thickness, (d–f) prepared from HBC-CN (3c) monolayers of different thickness. The thickness of the membranes is written at the top right corner of each image. The histograms in the insets show the distribution of the pore diameter with a step of 0.5 nm. The membranes are placed onto Quantifoil TEM grids type Multi A. Reprinted with permission from Ref. [6]. Copyright (2013) American Chemical Society.

Further studies demonstrated that the mechanical stiffness and the resonance frequency of a monolayer membrane can also be finely tuned via varying the precursor molecules. To investigate the mechanics of such ultrathin 2D sheets, mechanical characterization methods must be adapted to the nanoscale by combining them with nanoanalytical tools. Bulge testing was used to characterize the mechanical properties of freestanding films. The technique involves clamping of a freestanding membrane over an orifice and application of an overpressure to one side. The deflection of this 1-nm-thin carbon nanosheet was monitored by atomic force microscopy (AFM). The Young’s modulus and prestress were calculated from the obtained pressure–deflection relationship. Figure 14(a) shows a schematic diagram of the AFM bulge test setup. Freestanding CNMs with a thickness between 0.6 and 1.7 nm from diverse polyaromatic precursors were investigated by Zhang et al. [58, 59]. A correlation between the rigidity of the precursor molecules and the macroscopic mechanical stiffness of CNM was found. The data show that CNMs from rigid and condensed precursors like NPTH and pyrene thiols prove to exhibit higher Young’s moduli of 15−19 GPa, while CNMs from nonfused oligophenyls possess lower Young’s moduli of ~10 GPa. Figure 15 shows the elastic performance of CNM among other material classes. Materials with similar stiffness are porous ceramics and composites, whereas the class of polymers exhibits lower mechanical stability than the CNM.

Figure 14: (a) Schematic diagram of the AFM bulge test setup where a gas pressure is applied with compressed nitrogen, and the corresponding membrane deflection is recorded with an AFM. (b) Plot of Young’s moduli of CNMs prepared from different precursor molecules. (c) Residual stress of CNMs, where CNMs from bulky molecules are separately presented. Reprinted with permission from Ref. [59]. Copyright (2014) from American Chemical Society.
Figure 14:

(a) Schematic diagram of the AFM bulge test setup where a gas pressure is applied with compressed nitrogen, and the corresponding membrane deflection is recorded with an AFM. (b) Plot of Young’s moduli of CNMs prepared from different precursor molecules. (c) Residual stress of CNMs, where CNMs from bulky molecules are separately presented. Reprinted with permission from Ref. [59]. Copyright (2014) from American Chemical Society.

Figure 15: Young’s modulus plot of different materials [60].
Figure 15:

Young’s modulus plot of different materials [60].

The interpretation of Zhang et al.’s data relates the Young’s modulus to the properties of single molecules and to the structure of the pristine SAMs. The membranes with the highest stiffness are formed from molecules that exhibit higher molecular rigidity (less conformational freedom), whereas the terphenyl-based CNM is more mechanically stable than biphenylthiol-based CNM due to the denser packing of molecules within the SAM and to the larger number of cross-linking sites in the molecular skeleton. The degree of cross-linking of the SAM proved to be also an important parameter that has a strong impact on the membrane prestress or tensile stress. Although van der Waals interactions between the sheet and the substrate are the main factors that determine the constant tension of a self-supported monolayer, different prestresses for CNMs prepared from different precursor molecules, in spite of their similar thickness, were estimated. This implies that electron irradiation plays a key role, as a higher number of conformational degrees of freedom of the precursor molecules enable more ways to cross-link the molecules and thus achieve a lower prestress in the CNM, as in the case of TPT-CNMs. The ultimate tensile strength of CNM was also determined by means of bulge tests, measuring the ultimate pressure at which the membrane ruptures [58]. The tensile strength of biphenyl-based CNMs ranges from 400 to 700 MPa. In this range materials like graphite, iron and silicon nitride break. Of course, the tensile strength of CNM is orders of magnitude lower than that of graphene and carbon nanotubes, but compared to other nanomembranes, such as polymer membranes and the organic–inorganic IPNs, exhibiting a tensile strength of 105 MPa, the ultimate tensile strength of CNMs is five to six times higher.

Zhang et al. [61] also studied the mechanical motion of the CNM and showed that they can be applied as mechanical resonators. Their vibrational mode shapes were characterized and visualized by optical interferometry (Figure 16). The freestanding CNM was transferred onto Si substrates with square or rectangular orifices. A vibration of the membrane was actuated by applying a sinusoidal voltage to a piezoelectric disk on which the sample was glued. The dynamic behavior of the CNM’s vibrational modes is described by linear response theory of a membrane with negligible bending rigidity. A phase velocity of 61.6 ± 10 m/s and 55.06 ± 9 m/s was obtained for 4ʹ-nitro-1,1ʹ-biphenyl-4-thiol (NBPT)-based CNM and TPT–CNM, respectively. By comparing the dispersion relation to an analytical model, the static stress of the membranes of ~5 MPa was determined and found to be caused by the fabrication process. The extremely low mass and bending stiffness ensures that CNM resonators are in the membrane regime where the pretension of the CNM dominates over its elastic stiffness.

Figure 16: Several representative vibration modes of a 4ʹ-nitro-1,1ʹ-biphenylthiol CNM with dimensions of 50 × 50 μm2 at (a) 700 kHz, (b) 1,660 kHz, (c) 1,810 kHz. Reprinted with permission from Ref. [61]. Copyright (2015) from AIP Publishing LLC.
Figure 16:

Several representative vibration modes of a 4ʹ-nitro-1,1ʹ-biphenylthiol CNM with dimensions of 50 × 50 μm2 at (a) 700 kHz, (b) 1,660 kHz, (c) 1,810 kHz. Reprinted with permission from Ref. [61]. Copyright (2015) from AIP Publishing LLC.

Stimulated by the development of polymer and asymmetric hollow fiber membranes, membrane separation processes gained rapid progress not only in medical applications (artificial kidneys and lungs) and in the field of environmental protection (wastewater treatment, desalination, purification of flue gases), but even in energy renewable technologies. Since the end of the twentieth century, carbon materials, such as carbon molecular sieves, fullerenes, single and multiwall nanotubes, were extensively studied for their implementation in membrane technology. Two-dimensional nanomembranes with a thickness below ~5 nm and pores tuned to act as molecular sieves are predicted to be ideal separation membranes with many advantages over bulk membranes [62]. Especially, such extremely thin and dense carbon-based nanomembranes are considered to be effective and superior separation media for chemical and gas purification [63], offering the advantage of fast flow and lack of aging. Ai et al. [64] investigated the gas permeation performance of CNMs, built up from biphenylthiol (BPT) and p-nitrobiphenylthiol (NBPT) SAMs. Their permeance was measured onto a polydimethylsiloxane-supported membrane (PDMS-TFC, 150 nm PDMS/PAN/nonwoven) [65, 66] in order to enhance the mechanical stability and minimize roughness-induced strain. Considering the geometry and composition of the measured system – CNM/PDMS/PAN/nonwoven – it is intuitive to consider that the gas transport properties of PDMS determine and contribute significantly to the measured permeance values of CNM-PDMS. Therefore, it is necessary to consider the change of the permeance of bare PDMS due to the deposition of CNMs. A number of gases, such as hydrogen, helium, carbon dioxide, oxygen, nitrogen, argon, methane and ethane, have been tested and their corresponding permeance and ideal gas selectivity were obtained. The data are presented in relation to the penetrant kinetic diameters of the gases in Figure 17. They clearly show that the gas permeation of PDMS membranes reduces by the deposition of CNMs. This reduction is more pronounced for gas species with larger kinetic diameters. A simple model to extract the permeance of CNMs from the measured values of CNM-PDMS membranes revealed a different gas permeation behavior for single-layer and three-layer CNMs. In both cases, He and H2 showed substantially higher permeance values in comparison to gases with larger kinetic diameters with the exception of CO2, which displayed high permeances for single-layer CNMs and low values for three-layer CNMs. The established relationship of permeance and molecular size determines the molecular sieving mechanism of permeation through this ultrathin sheet. It can be understood by assuming that molecular-sized channels in the CNM dominate the permeation properties. These molecular-sized channels may form during the cross-linking process due to its statistical nature, which is expected to result in a random variation of intermolecular distances. Thus, even in a defect-free CNM, a high density of small free volumes (or pores) may form at locations of relatively large intermolecular distances, which act as the proposed molecular-sized channels.

Figure 17: (a) Permeance of the CNM part of the composite membranes that accounts for the finite permeance of the PDMS support as well as a partial coverage of the PDMS surface by CNM. round symbol – single layer of NBPT-CNM, square symbol – single layer of BPT-CNM, triangle, pointing up – three-layer BPT-CNM, triangle, pointing down – three-layer NBPT-CNM. (b) Schematic depiction of the proposed gas transport mechanism in single-layer CNMs. Molecular-sized channels (highlighted by bright regions) favor the permeation of CO2 and smaller gas molecules. (c) Schematic depiction of the proposed gas transport mechanism in multilayer CNMs. A low diffusion of CO2 in between the individual CNMs hinders its permeation through multilayer CNMs. Reprinted with permission from Ref. [64]. Copyright (2014) from Wiley-VCH.
Figure 17:

(a) Permeance of the CNM part of the composite membranes that accounts for the finite permeance of the PDMS support as well as a partial coverage of the PDMS surface by CNM. round symbol – single layer of NBPT-CNM, square symbol – single layer of BPT-CNM, triangle, pointing up – three-layer BPT-CNM, triangle, pointing down – three-layer NBPT-CNM. (b) Schematic depiction of the proposed gas transport mechanism in single-layer CNMs. Molecular-sized channels (highlighted by bright regions) favor the permeation of CO2 and smaller gas molecules. (c) Schematic depiction of the proposed gas transport mechanism in multilayer CNMs. A low diffusion of CO2 in between the individual CNMs hinders its permeation through multilayer CNMs. Reprinted with permission from Ref. [64]. Copyright (2014) from Wiley-VCH.

The transport through a three-layer CNM is consistent with lateral diffusion that is high for He and H2 but hindered for the larger gases including CO2. Lateral diffusion of gas molecules may be described by Knudsen-like diffusion or by condensation and surface flow of gas molecules in between the CNMs. Both possible mechanisms result in a permeance characteristic, which prefers the transport of smaller gases. The three-layer CNMs exhibit very small and constant values for all gases except the two gas species with the smallest kinetic diameters, which show more than one order of magnitude higher permeances. These higher permeance values cannot be explained by considering defects in the membrane as they would increase the permeance for all gas species that implies that this behavior is an intrinsic property of CNMs, resembling the permeation mechanism through glassy polymers with stiff polymer backbones, which also act like molecular sieves [67].

In addition, these experiments demonstrate that the gas permeation of PDMS membranes can be modified by the deposition of CNMs in a controlled way. The flexibility in the assembly of CNMs, by employing different precursor molecules as well as functionalizing CNMs, opens a path to molecular designed membranes also with respect to desired gas permeation characteristics.

This chapter describes a universal scheme to fabricate tailored CNM. CNMs of different size, thickness, elasticity and perforation have been generated. Their properties can be tailored by the choice of the precursor molecule, the structure of the SAM, the cross-linking and the transfer process. As the procedures that fabricate the CNMs are to a large extent compatible with nanofabrication protocols in industry, CNMs can be easily incorporated into electronic devices [68], into TEM grids as ultrathin carbon supports for high-contrast imaging [69], in separation technologies [64] or as protective coatings on chemically and mechanically sensitive technological components. This gives the material scientist a tool to produce membranes for technological applications. Here, a focus on material separation seems logical, as the manifold of diverse and specific separation problems is best addressed by a modular construction system of tailored CNMs. Beyond that the “similarity” of the CNM with biomembranes opens new opportunities for research. We can think of designing hybrid structures where biological membrane proteins are inserted into CNMs, creating an artificial 2D system with functional biological units. This could provide reference structures for the incorporation of single molecules into biotechnological sensors or devices. The future will show whether this will become a reality.

Acknowledgment

This article is also available in: Muellen, Feng, Chemistry of Carbon Nanostructures. De Gruyter (2016), isbn 978–3–11–028450–8.

Acknowledgments

The work reviewed in this chapter was financially supported by the Volkswagenstiftung, the Deutsche Forschungsgemeinschaft, the German Bundesministerium für Bildung und Forschung. The research leading to these results has also received funding from the European Union Seventh Framework Programme under grant agreement n°604391 Graphene Flagship.

References

[1] Turchanin A, Gölzhäuser A. Carbon nanomembranes from self-assembled monolayers: functional surfaces without bulk. Prog Surf Sci 2012;87(5–8):108–62.10.1016/j.progsurf.2012.05.001Search in Google Scholar

[2] Anselmetti D, Gölzhäuser A. Converting molecular monolayers into functional membranes. Angew Chem Int Ed 2014;53(46):12300–2.10.1002/anie.201406789Search in Google Scholar PubMed

[3] Rogers JA, Lagally MG, Nuzzo RG. Synthesis, assembly and applications of semiconductor nanomembranes. Nature 2011;477(7362):45–53.10.1038/nature10381Search in Google Scholar PubMed

[4] Wang EN, Karnik R. Water desalination: Graphene cleans up water. Nat Nano 2012;7(9):552–4.10.1038/nnano.2012.153Search in Google Scholar PubMed

[5] Geim AK, Novoselov KS. The rise of graphene. Nat Mater 2007;6(3):183–91.10.1142/9789814287005_0002Search in Google Scholar

[6] Angelova P, Vieker H, Weber N-E, Matei D, Reimer O, Meier I, et al. A universal scheme to convert aromatic molecular monolayers into functional carbon nanomembranes. ACS Nano 2013;7(8):6489–97.10.1021/nn402652fSearch in Google Scholar PubMed

[7] Cavallo F, Lagally MG. Semiconductors turn soft: inorganic nanomembranes. Soft Matter 2010;6(3):439–55.10.1039/B916582GSearch in Google Scholar

[8] Vendamme R, Onoue SY, Nakao A, Kunitake T. Robust free-standing nanomembranes of organic/inorganic interpenetrating networks. Nat Mater 2006;5(6):494–501.10.1038/nmat1655Search in Google Scholar PubMed

[9] (a) Yaghi OM, Li G, Li H. Selective binding and removal of guests in a microporous metal-organic framework. Nature 1995;378(6558):703–6; (b) Darbandi M, Arslan HK, Shekhah O, Bashir A, Birkner A, Wöll C. Fabrication of free-standing ultrathin films of porous metal-organic frameworks by liquid-phase epitaxy and subsequent delamination. Physica Status Solidi RRL 2010;4(8–9):197–199.10.1038/378703a0Search in Google Scholar

[10] Moore TT, Mahajan R, Vu DQ, Koros WJ. Hybrid membrane materials comprising organic polymers with rigid dispersed phases. AIChE J 2004;50(2):311–21.10.1002/aic.10029Search in Google Scholar

[11] (a) Decher G. Fuzzy nanoassemblies: toward layered polymeric multicomposites. Science 1997;277(5330):1232–7; (b) Perez-Madrigal MM, Armelin E, Puiggali J, Aleman C. Insulating and semiconducting polymeric free-standing nanomembranes with biomedical applications. J Mater Chem B 2015;3(29):5904–32; (c) Sakamoto J, van Heijst J, Lukin O, Schlüter AD. Two-dimensional polymers: just a dream of synthetic chemists? Angew Chem Int Ed 2009;48(6):1030–69.10.1126/science.277.5330.1232Search in Google Scholar

[12] Karan S, Jiang Z, Livingston AG. Sub-10 nm polyamide nanofilms with ultrafast solvent transport for molecular separation. Science 2015; 348(6241):1347–51.10.1126/science.aaa5058Search in Google Scholar PubMed

[13] Lvov Y, Decher G, Möhwald H. Assembly, structural characterization, and thermal behavior of layer-by-layer deposited ultrathin films of poly(vinylsulfate) and poly(allylamine). Langmuir 1993;9(2):481–86.10.1021/la00026a020Search in Google Scholar

[14] Raaijmakers MJ, Hempenius MA, Schön PM, Vancso GJ, Nijmeijer A, Wessling M, et al. Sieving of hot gases by hyper-cross-linked nanoscale-hybrid membranes. J Am Chem Soc 2014;136(1):330–5.10.1021/ja410047uSearch in Google Scholar PubMed

[15] Novoselov KS, Jiang D, Schedin F, Booth TJ, Khotkevich VV, Morozov SV, et al. Two-dimensional atomic crystals. Proc Natl Acad Sci USA 2005;102(30):10451–53.10.1073/pnas.0502848102Search in Google Scholar PubMed PubMed Central

[16] Novoselov KS, Fal’ko VI, Colombo L, Gellert PR, Schwab MG, Kim KG. A roadmap for graphene. Nature 2012;490:192–200.10.1038/nature11458Search in Google Scholar PubMed

[17] Colson JW, Woll AR, Mukherjee A, Levendorf MP, Spitler EL, Shields VB, et al. Oriented 2D covalent organic framework thin films on single-layer graphene. Science 2011;332(6026):228–31.10.1126/science.1202747Search in Google Scholar PubMed

[18] Schultz MJ, Zhang XY, Unarunotai S, Khang DY, Cao Q, Wang CJ, et al. Synthesis of linked carbon monolayers: films, balloons, tubes, and pleated sheets. P Natl Acad Sci USA 2008;105(21):7353–8.10.1073/pnas.0710081105Search in Google Scholar PubMed PubMed Central

[19] (a) Grill L, Dyer M, Lafferentz L, Persson M, Peters MV, Hecht S. Nano-architectures by covalent assembly of molecular building blocks. Nat Nanotechnol 2007;2(11):687–91; (b) Kissel P, Erni R, Schweizer WB, Rossell MD, King BT, Bauer T, et al. A two-dimensional polymer prepared by organic synthesis. Nat Chem 2012;4(4):287–91; (c) Bauer T, Zheng ZK, Renn A, Enning R, Stemmer A, Sakamoto J. et al. Synthesis of free-standing, monolayered organometallic sheets at the air/water interface. Angew Chem Int Ed 2011;50(34):7879–84; (d) Bally-Le Gall F, Friedmann C, Heinke L, Arslan H, Azucena C, Welle A. et al. free-standing nanomembranes based on selective CVD deposition of functional poly-p-xylylenes. ACS Nano 2015;9(2):1400–7; (e) Lackinger M. On-surface polymerization – a versatile synthetic route to two-dimensional polymers. Polym Int 2015;64(9):1073–8; (f) Schrettl S, Stefaniu C, Schwieger C, Pasche G, Oveisi E, Fontana Y. et al. Functional carbon nanosheets prepared from hexayne amphiphile monolayers at room temperature. Nat Chem 2014;6(6):468–76.10.1038/nnano.2007.346Search in Google Scholar PubMed

[20] Turchanin A, Beyer A, Nottbohm CT, Zhang X, Stosch R, Sologubenko AS, et al. One nanometer thin carbon nanosheets with tunable conductivity and stiffness. Adv Mater 2009;21:1233–7.10.1002/adma.200803078Search in Google Scholar

[21] Ward BW, Notte JA, Economou NP. Helium ion microscope: a new tool for nanoscale microscopy and metrology. J Vac Sci Technol B 2006;24(6):2871–4.10.1116/1.2357967Search in Google Scholar

[22] (a) Bell DC, Contrast mechanisms and image formation in helium ion microscopy. Micros Microanal 2009;15(2):147–53; (b) Beyer A, Vieker H, Klett R, Meyer zu Theenhausen H, Angelova P, Gölzhäuser A, Imaging of carbon nanomembranes with helium ion microscopy. Beilstein J Nanotechnol 2015;6:1712–20.10.1017/S1431927609090138Search in Google Scholar

[23] Geyer W, Stadler V, Eck W, Zharnikov M, Gölzhäuser A, Grunze M. Electron-induced crosslinking of aromatic self-assembled monolayers: negative resists for nanolithography. Appl Phys Lett 1999;75(16):2401–3.10.1063/1.125027Search in Google Scholar

[24] Turchanin A, El-Desawy M, Göelzhäuser A. High thermal stability of cross-linked aromatic self-assembled monolayers: Nanopatterning via selective thermal desorption. Appl Phys Lett 2007;90(5).10.1063/1.2437091Search in Google Scholar

[25] Turchanin A, Weber D, Büenfeld M, Kisielowski C, Fistul MV, Efetov KB, et al. Conversion of self-assembled monolayers into nanocrystalline Graphene: structure and electric transport. ACS Nano 2011;5. DOI: 10.1021/nn200297n10.1021/nn200297nSearch in Google Scholar

[26] (a) Zheng Z, Nottbohm CT, Turchanin A, Muzik H, Beyer A, Heilemann, M, et al. Janus nanomembranes: a generic platform for chemistry in two dimensions. Angew Chem Int Ed 2010;49(45):8493–7; (b) Schnietz M, Turchanin A, Nottbohm CT, Beyer A, Solak HH, Hinze P, et al. Chemically functionalized carbon nanosieves with 1-nm thickness. Small 2009;5(23): 2651–5; (c) Nottbohm CT, Turchanin A, Gölzhäuser A, Metallization of organic monolayers: electroless deposition of Cu onto cross-linked aromatic self-assembled monolayers. Z Phys Chem Int J Res Phys Chem Chem Phys 2008;222(5–6):917–26; (d) Völkel B, Kaltenpoth G, Handrea M, Sahre M, Nottbohm CT, Küller A, et al. Electrodeposition of copper and cobalt nanostructures using self-assembled monolayer templates. Surf Sci 2005;597(1–3):32–41; (e) Zheng Z, Zhang X, Neumann C, Emmrich D, Winter A, Vieker H, et al. Hybrid van der Waals heterostructures of zero-dimensional and two-dimensional materials. Nanoscale 2015;7(32):13393–7; (f) Beyer A, Turchanin A, Nottbohm CT, Mellech N, Schnietz M, Gölzhäuser A. Fabrication of metal patterns on freestanding graphenoid nanomembranes. J Vac Sci Technol B: Microelectron Nanometer Struct 2010;28(6):C6D5.10.1002/anie.201004053Search in Google Scholar

[27] Gölzhäuser A, Geyer W, Stadler V, Eck W, Grunze M, Edinger K, et al. Nanoscale patterning of self-assembled monolayers with electrons. J Vac Sci Technol B 2000;18(6):3414–18.10.1116/1.1319711Search in Google Scholar

[28] (a) Küller A, Eck W, Stadler V, Geyer W, Gölzhäuser A. Nanostructuring of silicon by electron-beam lithography of self-assembled hydroxybiphenyl monolayers. Appl Phys Lett 2003;82(21):3776–8; (b) Kaltenpoth G, Völkel B, Nottbohm CT, Gölzhäuser A, Buck M. Electrode modification by electron-induced patterning of self-assembled monolayers. J Vac Sci Technol B 2002;20(6):2734–8.10.1063/1.1578537Search in Google Scholar

[29] (a) Eck W, Stadler V, Geyer W, Zharnikov M, Gölzhäuser A, Grunze M. Generation of surface amino groups on aromatic self-assembled monolayers by low energy electron beams – A first step towards chemical lithography. Adv Mater 2000;12(11):805–8; (b) Chesneau F, Hamoudi H, Schuepbach B, Terfort A, Zharnikov M. Modification of self-assembled mono layers of perfluoroterphenyl-substituted alkanethiols by low-energy electrons. J Phys Chem C 2011;115(11):4773–82; (c) Frey S, Rong HT, Heister K, Yang YJ, Buck M, Zharnikov, M. Response of biphenyl-substituted alkanethiol self-assembled monolayers to electron irradiation: damage suppression and odd−even effects. Langmuir 2002;18(8):3142–50.10.1002/(SICI)1521-4095(200006)12:11<805::AID-ADMA805>3.0.CO;2-0Search in Google Scholar

[30] Eck W, Küller A, Grunze M, Völkel B, Gölzhäuser A. Freestanding nanosheets from crosslinked biphenyl self-assembled monolayers. Adv Mater 2005;17(21):2583–7.10.1002/adma.200500900Search in Google Scholar

[31] Turchanin A, Tinazli A, El-Desawy M, Grossann H, Schnietz M, Solak HH, et al. Molecular self-assembly, chemical lithography, and biochemical tweezers: a path for the fabrication of functional nanometer-scale protein arrays. Adv Mater 2008;20(3):471–+.10.1002/adma.200702189Search in Google Scholar

[32] Gölzhäuser A, Eck W, Geyer W, Stadler V, Weimann T, Hinze P, et al. Chemical nanolithography with electron beams. Adv Mater 2001;13(11):806–809.10.1002/1521-4095(200106)13:11%3C803::AID-ADMA806%3E3.0.CO;2-WSearch in Google Scholar

[33] Turchanin A, Käfer D, El-Desawy M, Wöll C, Witte G, Gölzhäuser A. Molecular mechanisms of electron-induced cross-linking in aromatic SAMs. Langmuir 2009;25(13):7342–52.10.1021/la803538zSearch in Google Scholar

[34] (a) Cyganik P, Vandeweert E, Postawa Z, Bastiaansen J, Vervaecke F, Lievens P, et al. Modification and stability of aromatic self-assembled monolayers upon irradiation with energetic particles. J Phys Chem B 2005;109(11):5085–4; (b) Tai Y, Shaporenko A, Eck W, Grunze M, Zharnikov M. Depth distribution of irradiation-induced cross-linking in aromatic self-assembled monolayers. Langmuir 2004;20(17):7166–0; (c) Weidner T, Ballav N, Grunze M, Terfort A, Zharnikov M. Modification of biphenylselenolate monolayers by low-energy electrons. Physica Status Solidi B Basic Solid State Phys 2009;246(7):1519–28.10.1021/jp0478209Search in Google Scholar

[35] Olsen C, Rowntree PA. Bond-selective dissociation of alkanethiol based self-assembled monolayers adsorbed on gold substrates, using low-energy electron beams. J Chem Phys 1998;108(9):3750–64.10.1063/1.475780Search in Google Scholar

[36] (a) Palmer RE, Rous PJ. Resonances in electron-scattering by molecules on surfaces. Rev Mod Phys 1992;64(2):383–440; (b) Illenberger E. Electron-attachment reactions in molecular clusters. Chem Rev 1992;92(7):1589–1609; (c) Illenberger E, Formation and evolution of negative ion resonances at surfaces. Surf Sci 2003;528(1–3):67–77.10.1103/RevModPhys.64.383Search in Google Scholar

[37] Silien C, Raeisaenen MT, Buck M, A Supramolecular Network as sacrificial mask for the generation of a nanopatterned binary self-assembled monolayer. Small 2010;6(3):391–4.10.1002/smll.200901909Search in Google Scholar

[38] Turchanin A, Schnietz M, El-Desawy M, Solak HH, David C, Gölzhäuser A. Fabrication of molecular nanotemplates in self-assembled monolayers by extreme-ultraviolet-induced chemical lithography. Small 2007;3(12):2114–9.10.1002/smll.200700516Search in Google Scholar

[39] Völkel B, Gölzhauser A, Müller HU, David C, Grunze M. Influence of secondary electrons in proximal probe lithography. J Vac Sci Technol B 1997;15(6):2877–81.10.1116/1.589748Search in Google Scholar

[40] Tobita S, Meinke M, Illenberger E, Christophorou LG, Baumgartel H, Leach S. Polycyclic aromatic-hydrocarbons – negative-ion formation following low-energy (0–15 eV) electron-impact. Chem Phys 1992;161(3):501–8.10.1016/0301-0104(92)80165-RSearch in Google Scholar

[41] Hüfner S. Photoelectron spectroscopy principals and application, 3rd ed. Berlin: Springer, 2003.10.1007/978-3-662-09280-4Search in Google Scholar

[42] Laibinis PE, Whitesides GM, Allara DL, Tao YT, Parikh AN, Nuzzo RG. Comparison of the structures and wetting properties of self-assembled monolayers of n-alkanethiols on the coinage metal surfaces, copper, silver, and gold. J Am Chem Soc 1991;113(19):7152–67.10.1021/ja00019a011Search in Google Scholar

[43] Heister K, Zharnikov M, Grunze M, Johansson LSO, Ulman A. Characterization of X-ray induced damage in alkanethiolate monolayers by high-resolution photoelectron spectroscopy. Langmuir 2001;17(1):8–11.10.1021/la001101dSearch in Google Scholar

[44] Woodruff DP. The interface structure of n-alkylthiolate self-assembled monolayers on coinage metal surfaces. Phys Chem Chem Phys 2008;10(48):7211–7221.10.1039/b813948bSearch in Google Scholar PubMed

[45] (a) Edinger K, Grunze M, Wöll C. Corrosion of gold by alkane thiols. Berichte Der Bunsen-Gesellschaft-Phys Chem Chem Phys 1997;101(12):1811–15; (b) Yang GH, Liu GY, New insights for self-assembled monolayers of organothiols on Au(111) revealed by scanning tunneling microscopy. J Phys Chem B 2003;107(34):8746–59.10.1002/bbpc.19971011204Search in Google Scholar

[46] (a) Maksymovych P, Voznyy O, Dougherty DB, Sorescu DC, Yates JT, Jr. Gold adatom as a key structural component in self-assembled monolayers of organosulfur molecules on Au(111). Prog Surf Sci 2010;85(5–8):206–40; (b) Maksymovych P, Sorescu DC, Yates JT, Jr. Gold-adatom-mediated bonding in self-assembled short-chain alkanethiolate species on the Au(111) surface. Phys Rev Lett 2006;97(14) 146103.10.1016/j.progsurf.2010.05.001Search in Google Scholar

[47] Cossaro A, Mazzarello R, Rousseau R, Casalis L, Verdini A, Kohlmeyer A, et al. X-ray diffraction and computation yield the structure of alkanethiols on gold(111). Science 2008;321(5891):943–946.10.1126/science.1158532Search in Google Scholar PubMed

[48] Matei DG, Muzik H, Gölzhäuser A, Turchanin A. Structural investigation of 1,1’-biphenyl-4-thiol self-assembled monolayers on Au(111) by scanning tunneling microscopy and low-energy electron diffraction. Langmuir 2012;28(39):13905–11.10.1021/la803538zSearch in Google Scholar PubMed

[49] Azzam W, Cyganik P, Witte G, Buck M, Wöll C. Pronounced odd–even changes in the molecular arrangement and packing density of biphenyl-based thiol SAMs: A combined STM and LEED study. Langmuir 2003;19(20):8262–70.10.1021/la302821wSearch in Google Scholar PubMed

[50] (a) Heimel G, Romaner L, Bredas J-L, Zojer E, Odd–even effects in self-assembled monolayers of omega-(biphenyl-4-yl)alkanethiols: A first-principles study. Langmuir 2008;24(2): 474–82; (b) Heimel G, Romaner L, Zojer E, Bredas J-L, Toward control of the metal-organic interfacial electronic structure in molecular electronics: a first-principles study on self-assembled monolayers of pi-conjugated molecules on noble metals. Nano Lett 2007;7(4):932–40.10.1021/la7023814Search in Google Scholar PubMed

[51] Charbonneau GP, Delugeard Y. Structural transition in polyphenyls. 5. Biphenyl – 3-Dimensional data and new refinement at 293 K. Acta Crystallogr Sect B-Struct Sci 1977;33(MAY13):1586–8.10.1107/S0567740877006566Search in Google Scholar

[52] Matei DG, Weber, N-E, Kurasch S, Wundrack S, Woszczyna M, Grothe M, et al. Functional single-layer graphene sheets from aromatic monolayers. Adv Mater 2013;25(30):4146–51.10.1002/adma.201300651Search in Google Scholar PubMed

[53] (a) Feng X, Pisula W, Müllen K. Large polycyclic aromatic hydrocarbons: synthesis and discotic organization. Pure Appl Chem 2009;81(12):2203–24; (b) Wagner C, Kasemann D, Golnik C, Forker R, Esslinger M, Müllen K, et al. Repulsion between molecules on a metal: Monolayers and submonolayers of hexa-peri-hexabenzocoronene on Au(111). Phys Rev B 2010; 81(3):035423.10.1351/PAC-CON-09-07-07Search in Google Scholar

[54] Castner DG, Hinds K, Grainger DW. X-ray photoelectron spectroscopy sulfur 2p study of organic thiol and disulfide binding interactions with gold surfaces. Langmuir 1996;12(21):5083–86.10.1021/la960465wSearch in Google Scholar

[55] Zharnikov M, Grunze M. Spectroscopic characterization of thiol-derived self-assembling monolayers. J Phys-Condens Matter 2001;13(49):11333–65.10.1088/0953-8984/13/49/314Search in Google Scholar

[56] Kastler M, Pisula W, Wasserfallen D, Pakula T, Müllen K. Influence of alkyl substituents on the solution- and surface-organization of hexa-peri-hexabenzocoronenes. J Am Chem Soc 2005;127(12):4286–96.10.1021/ja0430696Search in Google Scholar PubMed

[57] Vericat C, Vela ME, Benitez G, Carro P, Salvarezza RC. Self-assembled monolayers of thiols and dithiols on gold: new challenges for a well-known system. Chem Soc Rev 2010;39(5):1805–34.10.1039/b907301aSearch in Google Scholar PubMed

[58] Zhang X, Beyer A, Gölzhäuser A Mechanical characterization of carbon nanomembranes from self-assembled monolayers. Beilstein J Nanotechnol 2011;2:826–33.10.3762/bjnano.2.92Search in Google Scholar PubMed PubMed Central

[59] Zhang X, Neumann C, Angelova P, Beyer A, Gölzhäuser A. Tailoring the mechanics of ultrathin carbon nanomembranes by molecular design. Langmuir 2014;30(27):8221–7.10.1021/la501961dSearch in Google Scholar PubMed

[60] Lovatt AM, Publishers S, Shercliff HR, Withers PJ. Material selection and processing P. Southgate Publishers, 2000 http://www-materials.eng.cam.ac.uk/mpsite/physics/introduction.Search in Google Scholar

[61] Zhang X, Waitz R, Yang F, Lutz C, Angelova P, Gölzhäuser A, et al. Vibrational modes of ultrathin carbon nanomembrane mechanical resonators. Appl Phys Lett 2015;106(6):063107.10.1063/1.4908058Search in Google Scholar

[62] Gozde O-I, Anna Maria C, Karen KG. CVD of polymeric thin films: applications in sensors, biotechnology, microelectronics/organic electronics, microfluidics, MEMS, composites and membranes. Rep Prog Phys 2012;75(1):016501.10.1088/0034-4885/75/1/016501Search in Google Scholar PubMed

[63] Koenig SP, Wang L, Pellegrino J, Bunch JS. Selective molecular sieving through porous graphene. Nat Nano 2012;7(11):728–32.10.1038/nnano.2012.162Search in Google Scholar PubMed

[64] Ai M, Shishatskiy S, Wind J, Zhang X, Nottbohm CT, Mellech N, et al. Carbon nanomembranes (CNMs) supported by polymer: mechanics and gas permeation. Adv Mater 2014;26(21):3421–6.10.1002/adma.201304536Search in Google Scholar PubMed

[65] Wilfredo Y, Anja C, Jan W, Klaus-Viktor P. Nanometric thin film membranes manufactured on square meter scale: ultra-thin films for CO2 capture. Nanotechnology 2010;21(39): 395301.10.1088/0957-4484/21/39/395301Search in Google Scholar PubMed

[66] D. Behling, K. Hattenbach, K. Ohlrogge, K.-V. Peinemann, J. Wind, U.S. Patent 4,994,094, 1991.Search in Google Scholar

[67] S. P. Nunes, K.-V. Peinemann, in Membrane Technology in the Chemical Industry, (Eds: S. P. Nunes, K.-V. Peinemann), Wiley-VCH, Weinheim, Germany 2001, Ch. 1.10.1002/3527600388Search in Google Scholar

[68] Woszczyna M, Winter A, Grothe M, Willunat A, Wundrack S, Stosch R, et al. All-carbon vertical van der Waals heterostructures: non-destructive functionalization of Graphene for electronic applications. Adv Mater 2014;26(28):4831–7.10.1002/adma.201400948Search in Google Scholar PubMed

[69] (a) Nottbohm CT, Beyer A, Sologubenko AS, Ennen I, Hutten A, Rosner H, et al. Novel carbon nanosheets as support for ultrahigh-resolution structural analysis of nanoparticles. Ultramicroscopy 2008;108(9):885–92; (b) Rhinow D, Büenfeld M, Weber NE, Beyer A, Gölzhäuser A, Kühlbrandt W, et al. Energy-filtered transmission electron microscopy of biological samples on highly transparent carbon nanomembranes. Ultramicroscopy 2011;111(5):342–9.10.1016/j.ultramic.2008.02.008Search in Google Scholar PubMed

Published Online: 2017-3-18
Published in Print: 2017-3-1

© 2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 24.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/psr-2016-0105/html?lang=en
Scroll to top button