Surface-enhanced Raman spectroscopy of carbon nanomembranes

Biomolecular Photonics, Faculty of Physics, Bielefeld University, 33615 Bielefeld, Germany. Abstract. Surface-enhanced Raman scattering spectroscopy (...
0 downloads 0 Views 2MB Size
Article Cite This: Langmuir 2018, 34, 2692−2698

pubs.acs.org/Langmuir

Surface-Enhanced Raman Spectroscopy of Carbon Nanomembranes from Aromatic Self-Assembled Monolayers Xianghui Zhang,*,† Marcel Mainka,†,§ Florian Paneff,† Henning Hachmeister,‡ André Beyer,† Armin Gölzhaü ser,† and Thomas Huser*,‡ †

Physics of Supramolecular Systems and Surfaces, Faculty of Physics, and ‡Biomolecular Photonics, Faculty of Physics, Bielefeld University, 33615 Bielefeld, Germany S Supporting Information *

ABSTRACT: Surface-enhanced Raman scattering spectroscopy (SERS) was employed to investigate the formation of selfassembled monolayers (SAMs) of biphenylthiol, 4′-nitro-1,1′-biphenyl-4-thiol, and p-terphenylthiol on Au surfaces and their structural transformations into carbon nanomembranes (CNMs) induced by electron irradiation. The high sensitivity of SERS allows us to identify two types of Raman scattering in electron-irradiated SAMs: (1) Raman-active sites exhibit similar bands as those of pristine SAMs in the fingerprint spectral region, but with indications of an amorphization process and (2) Ramaninactive sites show almost no Raman-scattering signals, except a very weak and broad D band, indicating a lack of structural order but for the presence of graphitic domains. Statistical analysis showed that the ratio of the number of Raman-active sites to the total number of measurement sites decreases exponentially with increasing the electron irradiation dose. The maximum degree of cross-linking ranged from 97 to 99% for the three SAMs. Proof-of-concept experiments were conducted to demonstrate potential applications of Raman-inactive CNMs as a supporting membrane for Raman analysis.



electron-beam irradiation,7 whereas aromatic SAMs become mechanically stabilized via covalent cross-linking of adjacent molecules and turn into carbon nanosheets .8−10 The cross-linked aromatic SAMs, also known as carbon nanomembranes (CNMs), possess monomolecular thickness and represent a new type of functional quasi-two-dimensional material.11 A variety of aromatic molecules have been used to prepare CNMs with well-defined thicknesses ranging from 0.6 to 3 nm.12 As the CNMs are derived from the precursor molecules, there is a correlation between the rigidity of the precursor molecules and the mechanical stiffness of corresponding CNMs, which has been revealed by bulge testing in atomic force microscopy.13,14 CNMs can be utilized as supporting films for transmission electron microscopy (TEM),15,16 as ultrathin membranes for gas separations,17 and potentially as thin dielectric layers for molecular electronics.18,19 The outstanding mechanical properties also

INTRODUCTION Carbon possesses the most unique bonding properties, which enable it to form different allotropes, such as diamond, fullerene and carbon nanotubes, graphite, and graphene. Because of the different anisotropy of sp2 and sp3 hybridized orbitals, graphite exhibits very different physical properties compared to diamond. It is intuitive to bridge these two allotropes by preparing new carbon materials with mixed sp2 and sp3 phases with controlled spatial distributions. Indeed, there are a variety of carbon materials consisting of both sp2 and sp3 phases, such as tetrahedral amorphous carbon (or diamond-like carbon) and nanocrystalline diamond.1,2 Local and precise transformation of the sp3 hybridized carbon matrix into a sp2 phase can be achieved by utilizing an ion- or an electron-beam, and the properties of such a transformed carbon matrix have been investigated by several surface-sensitive analytical techniques.3,4 A bottom-up approach to fabricate carbon nanomaterials with mixed sp2 and sp3 phases is to form self-assembled monolayers (SAMs) on surfaces and then to polymerize them via ultraviolet (UV) irradiation.5,6 Aliphatic SAMs usually undergo fragmentation and structural disordering when being exposed to © 2018 American Chemical Society

Received: November 30, 2017 Revised: January 22, 2018 Published: January 29, 2018 2692

DOI: 10.1021/acs.langmuir.7b03956 Langmuir 2018, 34, 2692−2698

Article

Langmuir

that, the BPT−CNM/PMMA stack was transferred to a pure water bath and subsequently fished out with a Si3N4/Si substrate. The sample was immersed into acetone for ∼45 min to dissolve the PMMA and blown dry in a stream of nitrogen. Second, 80 nm Au NPs were deposited on the BPT−CNM by the drop-casting method at room temperature. Again, two protective layers of PMMA were spin-coated on the Si/Si3N4/BPT−CNM/Au-NPs wafer. The sample was brought onto the surface of HF (48%) for 5 s to weaken the adhesive forces between the wafer and the CNM. Then, the system BPT−CNM/AuNPs/PMMA was separated from the Si3N4 and transferred onto a water surface. The BPT−CNM/Au-NPs/PMMA stack was fished out with a NBPT−SAM on an Au/mica substrate. Finally, the PMMA was dissolved in acetone to achieve a system consisting of Mica/Au/ NBPT−SAM/BPT−CNM/Au-NPs. Raman Spectroscopy. Raman measurements were conducted using a custom-built Raman microspectroscopy system using a 785 nm wavelength semiconductor laser (Innovative Photonic Solutions) with a power of 10 mW in the focal point. For the acquisition of the spectra, we used a 60× objective with 1.2 NA (Olympus). For the testing of the sample surface for Raman activity (statistics), we used a 20× objective with 0.4 NA. A spectrometer with a 600 lines per millimeter grid (Acton SpectraPro 2300i) combined with a charge-coupled device (CCD) camera (Andor Newton) was used to acquire Raman spectra. The CCD camera was cooled to −75 °C, and full vertical binning was chosen to minimize noise. For testing sample surfaces for Raman activity, 260 sites with a 100 μm distance to the nearest neighbor were investigated using a micrometer screw-controlled translation stage to access the respective sites. By obtaining a minimum signal-to-noise ratio of 3 for the peaks at ∼1080 and ∼1600 cm−1, a site was assessed to be Raman-active. Because a minimum of 97% of the sites showed Raman activity for a pristine BPT−SAM, the maximum random error is estimated to be ∼3% of all measured sites. Helium Ion Microscopy (HIM). A Carl Zeiss ORION PLUS microscope was utilized to conduct HIM measurements. For the helium ion beam, we used energies of 33−38 keV at beam currents of 0.2−1.4 pA. A 10 μm aperture was used for all measurements. For the detection of secondary electrons, an Everhart−Thornley detector with a grid voltage of 500 V was employed. The working distance ranged from 10 to 29 mm.

make them suitable candidates for nanoelectromechanical systems.20 The conversion of aromatic SAMs into CNMs has been investigated by using different complementary surface spectroscopic techniques and theoretical methods.21−25 Surface-enhanced Raman scattering spectroscopy (SERS) is a powerful surface-sensitive analytical technique that allows for detecting chemical and structural information of molecules on metallic surfaces,26−28 as well as on other SERS-active substrates.29−31 The dominant contribution of the SERS effect arises from the localized excitation of surface plasmons, which leads to an electromagnetic field enhancement at nanoscopic surface features termed as hot spots.32 There is usually another enhancement mechanism called chemical enhancement because of the formation of charge−charge complexes.33 SERS has been used to characterize the bonding, structure, and orientation of adsorbed molecules at Au and Ag surfaces.34−37 On the other hand, SAMs have been employed as model systems to study the influence of nanoparticle (NP) shapes and their optical properties on the SERS enhancement factors.38 Structural transformation of both aliphatic SAMs and fullerene molecules into amorphous carbon by high-energy electrons has been recently investigated using SERS.39 In this paper, we use biphenylthiol (BPT), 4′-nitro-1,1′biphenyl-4-thiol (NBPT), and p-terphenylthiol (TPT) as model molecules to form SAMs on Au(111) and prepare CNMs via electron irradiation. SERS assisted by Au NPs is employed not only to confirm the chemisorption of precursor molecules on Au surfaces but also to monitor structural changes involved in the conversion of pristine SAMs to cross-linked monolayers induced by electron-beam irradiation. Furthermore, the statistics are used to estimate the cross-linking degree in the three SAMs from their Raman-scattering spectra. Finally, we present a proof of concept of how CNMs could be used as an ultrathin supporting layer for Raman analysis.





EXPERIMENTAL SECTION

Molecules and Materials. BPT was bought from Sigma-Aldrich. NBPT was purchased from Taros Chemicals (Dortmund, Germany). P-Terphenylthiol was specially synthesized. Thermally evaporated Au films (300 nm) on mica supports (Georg Albert physical vapor deposition coatings) were used as substrates for the SAM preparation. Suspensions of Au NPs in water were acquired from G. Kisker GbR (Steinfurt, Germany). Preparation of SAMs and CNMs. Cleaning of the Au surfaces was achieved with an UV/ozone-cleaner (FHR, UVOH 150 LAB) and subsequent immersing in absolute ethanol. The SAMs were prepared by immersion of the cleaned Au substrate into ∼10 mM solutions of the precursor molecules in dried and degassed N,N-dimethylformamide under a nitrogen atmosphere in a sealed flask for 72 h. Crosslinking of the SAMs was conducted by exposing the SAMs to 100 eV electrons (SPECS flood gun) in a high vacuum (10−7 mbar) using doses of 5, 10, 20, and 60 mC·cm−2. The Au NP suspension was deposited as-received onto SAMs or CNMs by the drop-casting method at room temperature. Preparation of a BPT−CNM Carrying 80 nm Au NPs and Transfer it onto a NBPT−SAM. To achieve a system consisting of Au/NBPT−SAM/BPT−CNM/Au−NPs, a BPT−CNM was first transferred onto a Si3N4/Si substrate. To this end, two layers of poly(methyl methacrylate) (PMMA; first layer: AR-P 631.04 and second layer: AR-P 671.04) with an overall thickness of ∼400 nm, were spin-cast on the sample and baked on a hot plate at 90 °C for 5 min. The underlying mica support was separated from the Au/BPT− CNM/PMMA structure by a slight dipping into water of one of the edges/corners of the sample. The separated Au/BPT−CNM/PMMA structure was transferred from the water surface to an I2/KI/H2O etching bath (1:4:10) where the Au layer would be dissolved. After

RESULTS AND DISCUSSION Figure 1a shows a schematic illustration of a pristine BPT− SAM on a gold surface. When the SAM is exposed to lowenergy electrons, the precursor molecules are dehydrogenated and cross-linked. For SERS measurements, we adopted a sphere-plane configuration by depositing Au NPs on pristine and cross-linked monolayers on Au(111) substrates, as schematically shown in Figure 1b.

Figure 1. (a) Formation of a biphenylthiol SAM (BPT−SAM) on the Au substrate. Low-energy electron irradiation leads to the cross-linking of molecules and the conversion to a CNM. (b) Deposition of Au NPs on a SAM or a CNM for SERS analysis. 2693

DOI: 10.1021/acs.langmuir.7b03956 Langmuir 2018, 34, 2692−2698

Article

Langmuir

Figure 2. Normal Raman spectra (upper plots) of precursor molecules in powder form: (a) BPT, (b) NBPT, and (c) TPT. The SER spectra (lower plots) of their corresponding SAMs on Au(111) substrates, where 80 nm Au NPs were used.

at 1339 cm−1 does not show any shifts in the SER spectrum.44 The in-plane rocking mode (b1 vibration) of the NO2 group at 530 cm−1 becomes stronger in the SER spectrum. The in-plane bending mode (a1 vibration) of the NO2 group at 856 cm−1 and the out-of-plane C−H angle deformation at 820 cm−1 (a2 fundamental, ν11) can be found in both spectra.45 The band at 763 cm−1 assigned to be the out-of-plane C−H bending with a contribution from the C−S−H bending mode becomes absent in the SAM.35,40 The C−C−C angle deformation at 625 cm−1 (b1 fundamental, ν18) is absent in the SER spectrum.45 The C− C−C angle deformation at 407 cm−1 (a1 fundamental, ν18′) and the NO2 out-of-plane rocking vibration at 420 cm−1 (b2 fundamental), which are resolved in the Raman spectrum of NBPT molecules, merge into one peak in the SER spectrum.45 NBPT molecules possess a symmetric ring breathing mode at 1110 cm−1. However, there was a pair of bands at 1081 and 1114 cm−1 in the SER spectrum of the NBPT−SAM. The same spectral changes were also found in the p-nitrothiophenol adsorbed on a silver surface.44 No significant shifts of the interring C−C stretching mode at 1284 cm−1 and of the in-plane deformation mode of the C−H bond at 1195 cm−1 were found in the SER spectrum of the NBPT−SAM. The ring mode at 1591 cm−1 appeared to have a downshift of 7 cm−1 and a broadening of the peak. These changes also suggest the chemisorption of NBPT molecules on the gold surface. The Raman spectrum of the TPT molecules and the SER spectrum of the TPT−SAM are also compared in Figure 2c. The splitting of the band at 1592 cm−1 into bands of 1584 and 1601 cm−1 is pronounced, indicating a downshifting of the mode associated with lower phenyl rings because of the chemical adsorption of TPT molecules.46 The band at 1272 cm−1 in the spectrum of TPT molecules, which is assigned to the inter-ring C−C stretching mode, exhibits an upshift of 3 cm−1 in the spectrum of the TPT−SAM.46 The in-plane deformation mode of the C−H bond at 1223 cm−1 is downshifted by ∼4 cm−1. The symmetric breathing mode of the phenyl ring at 1107 cm−1 exhibits a downshift by ∼26 cm−1 and an increase in intensity in the SER spectrum. The out-ofplane deformation of the C−H bond and the C−C bond at 1040 cm−1 appears weaker in the SER spectrum. The in-plane

Normal Raman spectra of three precursor molecules in powder form have been compared with the SER spectra of their corresponding SAMs formed on Au(111) substrates. In the Raman spectrum (the upper plot of Figure 2a) of BPT molecules, the bands at 2560 and 913 cm−1 are assigned to the stretching mode of the S−H bond and the deformation mode of the C−S−H vibration, respectively. The former band is not accessible by SERS, and the latter one is absent in the SER spectrum (the lower plot of Figure 2a). The band at 759 cm−1 assigned to be the out-of-plane C−H bending with a contribution from the C−S−H bending mode becomes significantly depressed in the SAM.35,40 Note that previous assignments of the band at 270 cm−1 to the Au−S stretching mode can be excluded because of the lack of Au surfaces here.35,41 The bands at 271 and 320 cm−1 can be probably assigned to the out-of-plane shearing motion (B2g fundamental) and the in-plane ring−ring shearing motion (Ag fundamental), respectively.40,42 The disappearance of both bands with the emergence of a new band at 286 cm−1 occurs in the SER spectrum of the BPT−SAM. In the SER spectrum, a new band appears at 479 cm−1 and is assigned to an out-of-plane phenyl vibration coupled with a torsion mode around the C−S bond.43 No significant shifts of the inter-ring C−C stretching mode at 1282 cm−1 were found in the SER spectrum of the BPT−SAM. The symmetric breathing mode of the phenyl ring that appeared at 1105 cm−1 in the bulk phase had a downshift of ∼26 cm−1 and a substantial increase in intensity in the SER spectrum. An apparent Fermi resonance doublet that occurs at about 1595 cm−1 in the bulk phase is found at 1584 and 1601 cm−1 in the SER spectrum, respectively. The significant downshifts of ring modes and the broadening of bands are characteristic of the chemisorbed SAM.34,35 All these changes indicate the chemisorption of BPT molecules to the Au surface and the absence of physisorbed molecules on the surface. Detailed assignment of the Raman bands is given in Table S1. The Raman spectrum of NBPT molecules and the SER spectrum of the NBPT−SAM are shown in Figure 2b. In the Raman spectrum of NBPT molecules in powder form, the band at 2561 cm−1 corresponds to the stretching vibration of the S− H bond. The symmetric stretching vibration of the NO2 group 2694

DOI: 10.1021/acs.langmuir.7b03956 Langmuir 2018, 34, 2692−2698

Article

Langmuir

states in which biphenyl molecules have a flattened conformation.50 Tashiro et al. reported that twisted biphenyl structures in liquid-crystalline acrylate polymers also exhibited three characteristic Raman bands, that is, 290, 320, and 420 cm−1, and the band intensities decreased down to zero as the biphenyl group changed the conformation from a twisted form to a planar form.51 Therefore, the intensity of the band at 408 cm−1 could indicate the degree of the twisting biphenyl molecules in the system.50,51 The formation of a network might also require an interlinking of the lower phenyls along one direction and of the upper ones along a different direction.24 The twist angle of a BPT−SAM consisting of the CN functional group was determined to be 41°.52 However, the twist angle of a pristine BPT−SAM has not yet measured so far. As any perturbation of the electron delocalization upon electron irradiation would lead to higher twist angles between two phenyl rings, a slight increase in the intensity of the band at 408 cm−1 may be attributed to an enhanced dihedral twist as a result of the loss of aromaticity. The non-cross-linked molecules are surrounded by a network of cross-linked molecules and such a confinement may also have an effect on their molecular conformations. Moreover, as long-range structural vibrations spread throughout the entire cross-linked structure and the Raman bonds in the lattice-vibrational region of an amorphous biphenyl film were only observed below 120 cm−1,50 it is also possible that the cross-linking results in a strong downshift of these modes below the detectable range. In addition, Figure 3 shows that a Raman spectrum (in black) of a BPT−CNM (irradiated at 60 mC·cm−2) annealed at 520 °C exhibits a very similar D peak at 1355 cm−1 and an additional broad G peak at 1604 cm−1, which indicates an absolute absence of non-cross-linked molecules and also a presence of nanocrystalline graphitic domains. To explore the relationship between the degree of crosslinking and the corresponding Raman activity, we performed a statistic analysis on BPT−SAMs with different electron irradiation doses, that is, 5, 10, 20, and 60 mC·cm−2. For this purpose, 260 surface sites were tested for Raman activity for each sample, where Au NPs with a diameter of 80 nm were used. Figure 4 shows that the ratio of Raman-inactive sites to the total number of measurement sites increases exponentially with increasing irradiation dose, which correlates well with the degree of cross-linking: (1) for a BPT−SAM without any crosslinking, nearly all sites were found to be Raman-active; (2) for a partially cross-linked SAM with an irradiation dose of 20 mC· cm−2, the percentage of Raman-inactive sites increases rapidly

deformation of the C−H bond (the middle phenyl ring) at 1003 cm−1 appears much stronger in the SER spectrum, whereas the out-of-plane deformation of the C−C−H bond (the upper and lower phenyl rings) at 993 cm−1 appears largely suppressed. The intensity of the in-plane ring vibration mode at 778 cm−1 decreases dramatically. There are also a few new peaks below 750 cm−1 in the SER spectrum: (1) the out-ofplane deformation of the C−C bond at 233 cm−1; (2) the inplane deformation of the C−C bond at 370 and 404 cm−1; (3) the inter-ring bending mode of the C−C bonds at 503 and 523 cm−1; (4) the intra-ring bending mode of C−C−C bonds at 615 and 638 cm−1; and (5) the out-of-plane deformation of C− C bond at 720 cm−1.46 All these changes indicate the chemisorption of the TPT molecules and the formation of a well-ordered molecular monolayer on the Au surface. Figure 3 compares several selective SER spectra obtained from a pristine BPT−SAM, a BPT−CNM, and an annealed

Figure 3. Comparison of representative SER spectra taken on a pristine BPT−SAM (green) and on a BPT−CNM (blue: Ramanactive site, red: Raman-inactive site); the Raman spectrum of a BPT− CNM that was annealed at 520 °C (black).

BPT−CNM. We found two types of spectra from a BPT− CNM: the first spectrum (in blue) exhibits similar bands as those of a pristine SAM (in green), whereas the second one (in red) shows almost no Raman-scattering signal except a very broad D peak at 1360 cm−1. The D peak is due to the breathing modes of the phenyl rings, which arises from the preserved sp2 hybridized carbon.47 Actually, sp3 hybridized carbon centers should be generated at the expense of the loss of sp2 hybridized carbon because of the electron irradiation, as confirmed from the electron energy loss spectroscopy investigation.23,25 However, the Raman scattering cross section of the sp2 phase is about 50−230 times higher than that of sp3 phase.3,48 The SER spectrum of an irradiated SAM is very likely dominated by the non-cross-linked molecules, and any structural transformation may only appear as superimposed weak signals on that. Nevertheless, a slight increase in the relative intensities of two bands at 408 and 287 cm−1 can be noticed in the SER spectrum recorded on Raman-active sites. In the Raman spectrum of biphenyl, the band at 410 cm−1 was assigned to the deformation vibration which corresponds to two nearly degenerate fundamentals, that is, one of B3g symmetry and the other of Au symmetry.49 Ishii et al. reported that Raman bands at 273, 316, and 410 cm−1 were found in biphenyl films evaporated at 5 K in which biphenyl molecules twisted around the phenyl−phenyl bond.50 These characteristic bands disappear at temperatures above 140 K for the polycrystalline

Figure 4. Percentage of Raman-inactive spots increases drastically with increasing irradiation dose, where a fully cross-linked monolayer appears to be Raman-inactive. 2695

DOI: 10.1021/acs.langmuir.7b03956 Langmuir 2018, 34, 2692−2698

Article

Langmuir

Figure 5. (a) Au NPs were carried by a BPT−CNM and placed onto a pristine NBPT−SAM on Au. HIM micrograph showing the Au(111)/ NBPT−SAM/BPT−CNM/Au-NPs (from bottom to top) system. (b) SER spectrum (in red) obtained from the Au(111)/NBPT−SAM/BPT− CNM/Au-NPs, where it was overlaid with the spectrum (in gray) of the Au(111)/NBPT−SAM/Au-NPs. Inset: A schematic of 1 nm-thick CNM that behaves as a supporting layer for transferring Au NPs and as a well-defined spacer between the Au NPs and the Au film.

ultrathin and flexible organic films are needed in Raman spectroscopy, as the outstanding mechanical properties allow CNMs to be used as a freestanding support for microscopic objects such as Au dots56 and NPs.57 First, as a proof of concept, we deposited polystyrene (PS) beads with a diameter of 1.1 μm on freestanding CNMs supported by a copper grid. Raman signals from three PS beads are clearly distinguishable from the fluorescence signals of the CNM (see Figure S3a in the Supporting Information). Second, we employed a BPT− CNM to carry 80 nm Au NPs and transferred the stack onto a NBPT−SAM on Au(111). On one hand, Raman signals from individual BPT molecules should be absent because of the fact that non-cross-linked molecules cannot sustain a lift-off from their initial substrate during the transfer process. On the other hand, much lower intensities of Raman bands from the underlying NBPT−SAM were expected because the electric field enhancement factor decreases with increasing size of the interstices.58,59 Figure 5a,b shows a helium ion micrograph and the corresponding SER spectra of such an assembly. Even though the intensities of those characteristic bands become approximately one tenth of those of a pristine NBPT−SAM, the SER spectrum exhibits the same characteristic features of the underlying NBPT−SAM. These results prove the capability of CNMs to function as a flexible supporting material for Raman studies.

to ∼94%; and (3) for a fully cross-linked SAM with an irradiation dose of 60 mC·cm−2, the ratio of Raman-inactive sites remains constant at 97%. The same tendency was found in the SER spectra of the NBPT−SAM and the TPT−SAM. The ratio of Raman-inactive sites for both SAMs irradiated at 60 mC·cm−2 was ∼99%, which might be correlated with a slightly higher degree of conformational freedom of NBPT and TPT molecules. It has been reported that both NBPT−CNMs and TPT−CNMs possess lower residual stress of 4−5 MPa because of a more efficient cross-linking among adjacent molecules.14,20 These results support the hypothesis that a fully cross-linked CNM is Raman-inactive and only non-cross-linked molecules that reside in the hotspots contribute to the SER spectra. In addition, we also demonstrated the successful formation of BPT−SAMs and BPT−CNM shells on Au NPs via electron irradiation (see Figure S1 in the Supporting Information). The precise molecular structure of a cross-linked SAM remains unknown to some extent. It was proposed by CabreraSanfelix et al. based on density functional theory calculations that the dehydrogenated BPT molecules form “graphene-like” nanoflakes as building blocks.22 However, classical molecular dynamics investigations of the BPT−CNM suggest that the molecular system does not reach the ground state to form graphene fences but instead “freezes” into a metastable irregular configuration described by a local energy minimum.24 X-ray photoelectron spectroscopy has been employed to investigate the electron irradiation induced changes in a BPT−SAM.21 Although the C 1s peak position is known to depend on the aromaticity of the molecules, no peak shift but a slight in the intensity was observed, and only a reduction of C 1s intensity of the irradiated SAM down to ∼90% was observed there.21 Also, a complete absence of orientational long-range order in the irradiated NBPT−SAM was suggested by Meyerbroeker et al. based on the change in the C 1s K-edge near-edge X-ray absorption fine structure spectra.53 Because a transition from a metastable state to a more stable state can be readily achieved by thermal annealing, pristine amorphous CNMs can be crystallized with short-range structural order, where the existence of nm-scale graphene domains was revealed by TEM.54,55 Figure S2 shows another Raman spectrum of a BPT−CNM annealed at 930 °C that exhibits the D peak from a disordered sp2 phase and the in-plane bond-stretching vibrations of sp2 C−C bonds (the G peak at 1605 cm−1). The finding that 1 nm-thick CNMs are Raman-inactive potentially makes them promising for applications where



CONCLUSIONS In summary, we have investigated the formation of three aromatic SAMs on gold substrates and their conversion into CNMs upon electron irradiation by using SERS. Although the SERS effect is very sensitive to the non-cross-linked molecules, the amorphization process can be identified. By applying statistical analysis over the whole monolayer, we found that Raman-active sites decrease exponentially with the degree of cross-linking. The cross-linking induced disappearance of Raman bands arises from a lack of long-range structural order in CNMs. Proof-of-concept experiments were performed to use CNMs as a flexible Raman-inactive supporting material for Raman analysis.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.langmuir.7b03956. 2696

DOI: 10.1021/acs.langmuir.7b03956 Langmuir 2018, 34, 2692−2698

Article

Langmuir



Monolayers into Functional Carbon Nanomembranes. ACS Nano 2013, 7, 6489−6497. (13) Zhang, X.; Beyer, A.; Gölzhäuser, A. Mechanical Characterization of Carbon Nanomembranes from Self-Assembled Monolayers. Beilstein J. Nanotechnol. 2011, 2, 826−833. (14) Zhang, X.; Neumann, C.; Angelova, P.; Beyer, A.; Gölzhäuser, A. Tailoring the Mechanics of Ultrathin Carbon Nanomembranes by Molecular Design. Langmuir 2014, 30, 8221−8227. (15) Nottbohm, C. T.; Beyer, A.; Sologubenko, A. S.; Ennen, I.; Hütten, A.; Rösner, H.; Eck, W.; Mayer, J.; Gölzhäuser, A. Novel Carbon Nanosheets as Support for Ultrahigh-Resolution Structural Analysis of Nanoparticles. Ultramicroscopy 2009, 109, 379. (16) Rhinow, D.; Büenfeld, M.; Weber, N.-E.; Beyer, A.; Gölzhäuser, A.; Kühlbrandt, W.; Hampp, N.; Turchanin, A. Energy-Filtered Transmission Electron Microscopy of Biological Samples on Highly Transparent Carbon Nanomembranes. Ultramicroscopy 2011, 111, 342−349. (17) Ai, M.; Shishatskiy, S.; Wind, J.; Zhang, X.; Nottbohm, C. T.; Mellech, N.; Winter, A.; Vieker, H.; Qiu, J.; Dietz, K.-J.; Gölzhäuser, A.; Beyer, A. Carbon Nanomembranes (CNMs) Supported by Polymer: Mechanics and Gas Permeation. Adv. Mater. 2014, 26, 3421−3426. (18) Penner, P.; Zhang, X.; Marschewski, E.; Behler, F.; Angelova, P.; Beyer, A.; Christoffers, J.; Gölzhäuser, A. Charge Transport through Carbon Nanomembranes. J. Phys. Chem. C 2014, 118, 21687−21694. (19) Zhang, X.; Marschewski, E.; Penner, P.; Beyer, A.; Gölzhäuser, A. Investigation of Electronic Transport through Ultrathin Carbon Nanomembrane Junctions by Conductive Probe Atomic Force Microscopy and Eutectic Ga-In Top Contacts. J Appl Phys. 2017, 122, 055103. (20) Zhang, X.; Waitz, R.; Yang, F.; Lutz, C.; Angelova, P.; Gölzhäuser, A.; Scheer, E. Vibrational Modes of Ultrathin Carbon Nanomembrane Mechanical Resonators. Appl. Phys. Lett. 2015, 106, 063107. (21) Turchanin, A.; Käfer, D.; El-Desawy, M.; Wöll, C.; Witte, G.; Gölzhäuser, A. Molecular Mechanisms of Electron-Induced CrossLinking in Aromatic Sams. Langmuir 2009, 25, 7342−7352. (22) Cabrera-Sanfelix, P.; Arnau, A.; Sánchez-Portal, D. FirstPrinciples Investigation of Electron-Induced Cross-Linking of Aromatic Self-Assembled Monolayers on Au(111). Phys. Chem. Chem. Phys. 2010, 12, 1578−1584. (23) Amiaud, L.; Houplin, J.; Bourdier, M.; Humblot, V.; Azria, R.; Pradier, C.-M.; Lafosse, A. Low-Energy Electron Induced Resonant Loss of Aromaticity: Consequences on Cross-Linking in Terphenylthiol Sams. Phys. Chem. Chem. Phys. 2014, 16, 1050−1059. (24) Mrugalla, A.; Schnack, J. Classical Molecular Dynamics Investigations of Biphenyl-Based Carbon Nanomembranes. Beilstein J. Nanotechnol. 2014, 5, 865−871. (25) Houplin, J.; Dablemont, C.; Sala, L.; Lafosse, A.; Amiaud, L. Electron Processing at 50 eV of Terphenylthiol Self-Assembled Monolayers: Contributions of Primary and Secondary Electrons. Langmuir 2015, 31, 13528−13534. (26) Nie, S. M.; Emery, S. R. Probing Single Molecules and Single Nanoparticles by Surface-Enhanced Raman Scattering. Science 1997, 275, 1102−1106. (27) Schlücker, S. Surface-Enhanced Raman Spectroscopy: Concepts and Chemical Applications. Angew. Chem., Int. Ed. 2014, 53, 4756− 4795. (28) Gruenke, N. L.; Cardinal, M. F.; McAnally, M. O.; Frontiera, R. R.; Schatz, G. C.; Van Duyne, R. P. Ultrafast and Nonlinear SurfaceEnhanced Raman Spectroscopy. Chem. Soc. Rev. 2016, 45, 2263−2290. (29) Toccafondi, C.; Zaccaria, R. P.; Dante, S.; Salerno, M. Fabrication of Gold-Coated Ultra-Thin Anodic Porous Alumina Substrates for Augmented SERS. Materials 2016, 9, 403. (30) Salerno, M.; Shayganpour, A.; Salis, B.; Dante, S. SurfaceEnhanced Raman Scattering of Self-Assembled Thiol Monolayers and Supported Lipid Membranes on Thin Anodic Porous Alumina. Beilstein J. Nanotechnol. 2017, 8, 74−81.

Detailed assignment of Raman peaks; BPT−CNM on Au NPs; Raman spectrum of annealed CNMs; and Raman spectrum of polystyrene beads on CNM and on BK7 glass (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (X.Z.). *E-mail: [email protected] (T.H.). ORCID

Xianghui Zhang: 0000-0002-5544-5221 Thomas Huser: 0000-0003-2348-7416 Present Address §

M.M.: Weidmueller Group, 32758 Detmold, Germany.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge financial support from the German Ministry for Education and Research (BMBF) and Deutsche Forschungsgemeinschaft (SPP 1928). The authors thank Prof. Andreas Terfort for providing us with precursor molecules and Dr. Andreas Winter for the schematic pictures.



REFERENCES

(1) Robertson, J. Diamond-Like Amorphous Carbon. Mater. Sci. Eng., R 2002, 37, 129−281. (2) Williams, O. A. Nanocrystalline Diamond. Diamond Relat. Mater. 2011, 20, 621−640. (3) Ilie, A.; Durkan, C.; Milne, W. I.; Welland, M. E. Surface Enhanced Raman Spectroscopy as a Probe for Local Modification of Carbon Films. Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 66, 045412. (4) Ilie, A.; Ferrari, A. C.; Yagi, T.; Rodil, S. E.; Robertson, J.; Barborini, E.; Milani, P. Role of sp(2) Phase in Field Emission from Nanostructured Carbons. J. Appl. Phys. 2001, 90, 2024−2032. (5) Kim, T.; Crooks, R. M.; Tsen, M.; Sun, L. Polymeric SelfAssembled Monolayers .2. Synthesis and Characterization of SelfAssembled Polydiacetylene Mono- and Multilayers. J. Am. Chem. Soc. 1995, 117, 3963−3967. (6) Kim, T.; Chan, K. C.; Crooks, R. M. Polymeric Self-Assembled Monolayers .4. Chemical, Electrochemical, and Thermal Stability of ωFunctionalized, Self-Assembled Diacetylenic and Polydiacetylenic Monolayers. J. Am. Chem. Soc. 1997, 119, 189−193. (7) Zharnikov, M.; Grunze, M. Modification of Thiol-Derived SelfAssembling Monolayers by Electron and X-Ray Irradiation: Scientific and Lithographic Aspects. J. Vac. Sci. Technol., B: Microelectron. Nanometer Struct.–Process., Meas., Phenom. 2002, 20, 1793−1807. (8) Eck, W.; Küller, A.; Grunze, M.; Völkel, B.; Gölzhäuser, A. Freestanding Nanosheets from Crosslinked Biphenyl Self-Assembled Monolayers. Adv. Mater. 2005, 17, 2583. (9) Turchanin, A.; Schnietz, M.; El-Desawy, M.; Solak, H. H.; David, C.; Gölzhäuser, A. Fabrication of Molecular Nanotemplates in SelfAssembled Monolayers by Extreme-Ultraviolet-Induced Chemical Lithography. Small 2007, 3, 2114−2119. (10) Zhang, X.; Vieker, H.; Beyer, A.; Gölzhäuser, A. Fabrication of Carbon Nanomembranes by Helium Ion Beam Lithography. Beilstein J. Nanotechnol. 2014, 5, 188−194. (11) Turchanin, A.; Gölzhäuser, A. Carbon Nanomembranes. Adv. Mater. 2016, 28, 6075−6103. (12) Angelova, P.; Vieker, H.; Weber, N.-E.; Matei, D.; Reimer, O.; Meier, I.; Kurasch, S.; Biskupek, J.; Lorbach, D.; Wunderlich, K.; Chen, L.; Terfort, A.; Klapper, M.; Müllen, K.; Kaiser, U.; Gölzhäuser, A.; Turchanin, A. A Universal Scheme to Convert Aromatic Molecular 2697

DOI: 10.1021/acs.langmuir.7b03956 Langmuir 2018, 34, 2692−2698

Article

Langmuir (31) Takei, H.; Bessho, N.; Ishii, A.; Okamoto, T.; Beyer, A.; Vieker, H.; Gölzhäuser, A. Enhanced Infrared LSPR Sensitivity of Cap-Shaped Gold Nanoparticles Coupled to a Metallic Film. Langmuir 2014, 30, 2297−2305. (32) Schatz, G. C.; Young, M. A.; Van Duyne, R. P. Electromagnetic Mechanism of SERS. Top. Appl. Phys. 2006, 103, 19−45. (33) Jiang, X.; Campion, A. Chemical Effects in Surface-Enhanced Raman-Scattering:Pyridine Chemisorbed on Silver Adatoms on Rh(100). Chem. Phys. Lett. 1987, 140, 95−100. (34) Bandyopadhyay, K.; Vijayamohanan, K.; Venkataramanan, M.; Pradeep, T. Self-Assembled Monolayers of Small Aromatic Disulfide and Diselenide Molecules on Polycrystalline Gold Films: A Comparative Study of the Geometrical Constraint Using Temperature-Dependent Surface-Enhanced Raman Spectroscopy, X-Ray Photoelectron Spectroscopy, and Electrochemistry. Langmuir 1999, 15, 5314−5322. (35) Szafranski, C. A.; Tanner, W.; Laibinis, P. E.; Garrell, R. L. Surface-Enhanced Raman Spectroscopy of Aromatic Thiols and Disulfides on Gold Electrodes. Langmuir 1998, 14, 3570−3579. (36) Fleger, Y.; Mastai, Y.; Rosenbluh, M.; Dressler, D. H. Surface Enhanced Raman Spectroscopy of Aromatic Compounds on Silver Nanoclusters. Surf. Sci. 2009, 603, 788−793. (37) DeVetter, B. M.; Mukherjee, P.; Murphy, C. J.; Bhargava, R. Measuring Binding Kinetics of Aromatic Thiolated Molecules with Nanoparticles Via Surface-Enhanced Raman Spectroscopy. Nanoscale 2015, 7, 8766−8775. (38) Orendorff, C. J.; Gole, A.; Sau, T. K.; Murphy, C. J. SurfaceEnhanced Raman Spectroscopy of Self-Assembled Monolayers: Sandwich Architecture and Nanoparticle Shape Dependence. Anal. Chem. 2005, 77, 3261−3266. (39) Mojarad, N.; Tisserant, J.-N.; Beyer, H.; Dong, H.; Reissner, P. A.; Fedoryshyn, Y.; Stemmer, A. Monitoring the Transformation of Aliphatic and Fullerene Molecules by High-Energy Electrons Using Surface-Enhanced Raman Spectroscopy. Nanotechnology 2017, 28, 165701. (40) Katon, J. E.; Lippincott, E. R. The Vibrational Spectra and Geometrical Configuration of Biphenyl. Spectrochim. Acta 1959, 15, 627−650. (41) Carron, K. T.; Hurley, L. G. Axial and Azimuthal Angle Determination with Surface-Enhanced Raman-Spectroscopy:Thiophenol on Copper, Silver, and Gold Metal Surfaces. J. Phys. Chem. 1991, 95, 9979−9984. (42) Bree, A.; Zwarich, R.; Taliani, C. Raman Excitation Profiles for Some Ag-Modes of Biphenyl in the Pre-Resonance Region. Chem. Phys. 1982, 70, 257−267. (43) Dolamic, I.; Varnholt, B.; Bürgi, T. Far-Infrared Spectra of WellDefined Thiolate-Protected Gold Clusters. Phys. Chem. Chem. Phys. 2013, 15, 19561−19565. (44) Skadtchenko, B. O.; Aroca, R. Surface-Enhanced Raman Scattering of P-Nitrothiophenol:Molecular Vibrations of Its Silver Salt and the Surface Complex Formed on Silver Islands and Colloids. Spectrochim. Acta, Part A 2001, 57, 1009−1016. (45) Stephenson, C. V.; Coburn, W. C.; Wilcox, W. S. The Vibrational Spectra and Assignments of Nitrobenzene, Phenyl Isocyanate, Phenyl Isothiocyanate, Thionylaniline and Anisole. Spectrochim. Acta 1961, 17, 933−946. (46) daCosta, A. M. A.; Karger, N.; Amado, A. M.; Becucci, M. Raman Spectroscopic Study of Pure P-Terphenyl and Tetracene:PTerphenyl Doped Crystals. Solid State Ionics 1997, 97, 115−121. (47) Ferrari, A. C.; Robertson, J. Resonant Raman Spectroscopy of Disordered, Amorphous, and Diamondlike Carbon. Phys. Rev. B: Condens. Matter Mater. Phys. 2001, 64, 075414. (48) Sails, S. R.; Gardiner, D. J.; Bowden, M.; Savage, J.; Rodway, D. Monitoring the Quality of Diamond Films Using Raman Spectra Excited at 514.5 nm and 633 nm. Diamond Relat. Mater. 1996, 5, 589− 591. (49) Bree, A.; Zwarich, R. Torsionally Induced Bands in the RamanSpectrum of a Biphenyl Crystal. J. Raman Spectrosc. 1982, 12, 247− 250.

(50) Ishii, K.; Nakayama, H.; Tanabe, K.; Kawahara, M. Structural Relaxations in Amorphous Biphenyl. Raman-Spectroscopic Studies. Chem. Phys. Lett. 1992, 198, 236−240. (51) Tashiro, K.; Hou, J. A.; Kobayashi, M.; Inoue, T. X-Ray Structure Analyses and Vibrational Spectral Measurements of the Model Compounds of the Liquid-Crystalline Arylate Polymers. Finding out the Raman Bands Characteristic of the Twisted Biphenyl Structure in Association with the Solid-to-Liquid Crystalline PhaseTransition. J. Am. Chem. Soc. 1990, 112, 8273−8279. (52) Ballav, N.; Schüpbach, B.; Dethloff, O.; Feulner, P.; Terfort, A.; Zharnikov, M. Direct Probing Molecular Twist and Tilt in Aromatic Self-Assembled Monolayers. J. Am. Chem. Soc. 2007, 129, 15416. (53) Meyerbroeker, N.; Waske, P.; Zharnikov, M. Amino-Terminated Biphenylthiol Self-Assembled Monolayers as Highly Reactive Molecular Templates. J. Chem. Phys. 2015, 142, 101919. (54) Rhinow, D.; Weber, N.-E.; Turchanin, A. Atmospheric Pressure, Temperature-Induced Conversion of Organic Monolayers into Nanocrystalline Graphene. J. Phys. Chem. C 2012, 116, 12295−12303. (55) Turchanin, A.; Weber, D.; Büenfeld, M.; Kisielowski, C.; Fistul, M. V.; Efetov, K. B.; Weimann, T.; Stosch, R.; Mayer, J.; Gölzhäuser, A. Conversion of Self-Assembled Monolayers into Nanocrystalline Graphene: Structure and Electric Transport. ACS Nano 2011, 5, 3896−3904. (56) Beyer, A.; Turchanin, A.; Nottbohm, C. T.; Mellech, N.; Schnietz, M.; Gölzhäuser, A. Fabrication of Metal Patterns on Freestanding Graphenoid Nanomembranes. J. Vac. Sci. Technol., B: Microelectron. Nanometer Struct.–Process., Meas., Phenom. 2010, 28, C6d5−C6d10. (57) Zheng, Z.; Zhang, X.; Neumann, C.; Emmrich, D.; Winter, A.; Vieker, H.; Liu, W.; Lensen, M.; Gölzhäuser, A.; Turchanin, A. Hybrid Van Der Waals Heterostructures of Zero-Dimensional and TwoDimensional Materials. Nanoscale 2015, 7, 13393−13397. (58) Mubeen, S.; Zhang, S.; Kim, N.; Lee, S.; Krämer, S.; Xu, H.; Moskovits, M. Plasmonic Properties of Gold Nanoparticles Separated from a Gold Mirror by an Ultrathin Oxide. Nano Lett. 2012, 12, 2088− 2094. (59) Halas, N. J.; Moskovits, M. Surface-Enhanced Raman Spectroscopy: Substrates and Materials for Research and Applications. MRS Bull. 2013, 38, 607−611.

2698

DOI: 10.1021/acs.langmuir.7b03956 Langmuir 2018, 34, 2692−2698