1 Formation and Ripening of Self-Assembled Multilayers from the

of Chemistry, Stanford University, Stanford, CA 94305. 4Lam Research Corporation, Fremont,. California, USA 94538. ‡These authors contributed equall...
0 downloads 0 Views 935KB Size
Subscriber access provided by UOW Library

Article

Formation and Ripening of Self-Assembled Multilayers from the Vapor-Phase Deposition of Dodecanethiol on Copper Oxide David S. Bergsman, Tzu-Ling Liu, Richard G. Closser, Katie L Nardi, Nerissa Draeger, Dennis M. Hausmann, and Stacey F. Bent Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.8b02150 • Publication Date (Web): 26 Jul 2018 Downloaded from http://pubs.acs.org on July 31, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Formation and Ripening of Self-Assembled Multilayers from the Vapor-Phase Deposition of Dodecanethiol on Copper Oxide

David S. Bergsman‡1, Tzu-Ling Liu‡2, Richard G. Closser3, Katie L. Nardi4, Nerissa Draeger4, Dennis M. Hausmann4, and Stacey F. Bent1,* 1

Department of Chemical Engineering, 2Department of Materials Science and Engineering, 3Department

of Chemistry, Stanford University, Stanford, CA 94305. 4Lam Research Corporation, Fremont, California, USA 94538 ‡

These authors contributed equally to this work.

*[email protected]

Abstract The vapor-phase reaction of dodecanethiol (DDT) with copper oxide surfaces and the molecular level composition and structure of the resulting films were examined. Atomic force microscopy, cross-sectional transmission electron microscopy, and electron energy loss/electron dispersive spectroscopy reveal that instead of forming self-assembled monolayers, DDT etches CuO surfaces to create ~8 nm thick Cuthiolate multilayers. These layers are composed of surprisingly well-ordered crystallites, oriented either parallel or perpendicular to the substrate surface. Pre-etching of the CuO to expose the underlaying copper metal is shown to prevent the formation of multilayers and instead allow for the formation of the expected monolayers. Water contact angle and Fourier transform infrared spectroscopy are further shown to be ineffective at distinguishing the multilayer and monolayer thiol films. Interestingly, the multilayer films are unstable in air, ripening into particles 20 microns wide and several hundred nanometers tall over the course of a week. Air exposure also results in the slow oxidation of the sulfur and copper within the films at a rate similar to what has been seen before for DDT monolayers. As a result, the multilayers show no significant improvement over monolayers in the prevention of oxidation.

1 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 32

Introduction The liquid-phase self-assembly of alkanethiol monolayers onto copper has been studied over the past two decades for use in a variety of applications, such as the synthesis of nanostructures, the passivation of copper surfaces, the creation of chemical sensors, and the selective deposition of materials through atomic layer deposition (ALD)1–7. Although a majority of that effort has focused on adsorption on copper metal, the formation of thiol-based self-assembled monolayers (SAMs) on copper oxide is interesting in that thiols have been reported to etch and reduce the copper oxide surface4,5,16,8–15. In the case of cupric oxide (CuO), the mechanism behind this etching is generally believed to involve the oxidation of thiols to disulfide by the CuO, releasing two hydrogen atoms that combine with oxygen from the oxide surface to produce water. This process simultaneously reduces the copper oxide, which is eventually coated in an adsorbed thiolate SAM. In contrast, cuprous oxides (Cu2O) do not undergo thiol-based etching11 (except through the addition of other etchants, such as an alkaline solvent12) and instead thiols bond directly on the Cu2O to form SAMs..

Some reports suggested that the thiol-based etching of CuO may enable the formation of Cu-thiolate multilayers8,11–13,16–19. Some of those reports proposed that the etched CuO is incorporated into the thiol films as a Cu-thiol complex that can be many SAM layers thick. This multilayer is suspected to form either as a result of the inability of the Cu1+ ions to dissolve and be removed11, or because the Cu partially dissolves and then re-adsorbs onto the copper metal surface before complexing with thiol ligands17. Although the structure of those multilayers and the mechanism behind their formation were not well understood, pretreatment of the CuO with acids, such as acetic acic20,21, nitric acid14,22, or hydrochloric acid16,22,23, was found to remove the CuO layer, preventing the formation of multilayers during thiol deposition. To distinguish these multilayers from their monolayer counterparts, we use the term “selfassembled monolayer” or “SAM” only in reference to known, well-packed monolayer films.

2 ACS Paragon Plus Environment

Page 3 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

In recent years, the use of self-assembled monolayers to control area-selective ALD for back-end semiconductor processing has created renewed interest in the study of thiol deposition on copper, particularly through vapor-phase approaches that can be more easily incorporated into industrial semiconductor fabrication processes7,16,24. Self-assembled monolayers are also often used to prevent copper surface oxidation25,26 and corrosion23,27. However, only a handful of articles have studied the selfassembly of thiols on copper through the vapor-phase4,5,16,28, and fewer have reported the formation of Cuthiolate multilayers16.

Here, we examine the deposition of dodecanethiols (DDTs) on copper surfaces via a vapor-phase approach and address many of the above gaps in understanding. In this study, DDTs are exposed to copper pre-treated with UV-ozone and the resulting films are probed by a combination of microscopic, spectroscopic, and diffraction techniques. We show that the vapor-deposition of DDT results in wellaligned multilayer Cu-thiolate films when the copper substrate begins as CuO. We further show that this multilayer does not form when the copper oxide is removed prior to deposition, through treatment with acetic acid, similar to what has been seen previously. These multilayers are not observable with traditional SAM characterization techniques, like Fourier transform infrared (FTIR) spectroscopy and water contact angle (WCA) measurements, which show equal, if not better, WCAs and crystallinity in FTIR spectroscopy than do thiol monolayers.

Finally, the data presented here indicate that multilayers show similar, moderate passivation against oxidation when compared to traditional monolayers. Over the course of several days' exposure to air, the multilayers ripen into micron-scale particles. This ripening does not occur in a nitrogen-purged atmosphere. Air exposure also results in the general oxidation of the Cu and sulfur of the films into Cusulfonate. Overall, these results provide insight into the growth mechanism and properties of copperthiolate multilayers formed through the vapor-phase functionalization of copper oxide surfaces with thiols. 3 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 32

Experimental Methods Dodecanethiol (DDT, >98%) was purchased from Sigma-Aldrich and used without additional purification. Blanket copper (Cu) and copper-on-silicon oxide patterned substrates (Cu/SiO2) were provided by collaborators and created as follows. Blanket Cu wafers were prepared by sequentially depositing SiO2, TaN, and Cu onto a blanket Si wafer by chemical vapor deposition (SiO2) and physical vapor deposition (TaN, Cu). Patterned Cu/SiO2 substrates were prepared by conventional lithography, etching and thin film deposition processes, with the Cu deposited by electrochemical plating. Chemical mechanical planarization was also applied on the blank and patterned copper substrates.

Before DDT deposition, substrates were cleaned using 10 min of sonication in absolute ethanol (200 proof, purchased from Fisher Scientific) and 2 min of UV-ozone treatment in a Novascan PSD Series Digital UV Ozone System to remove organic contaminants from the surface and ensure the complete conversion of any surface copper oxide species to cupric oxide11. After the UV ozone cleaning, select substrates were immersed into glacial acetic acid (Macron, ACS grade) for 3 min to remove the native copper oxide and then immediately transferred into a home built vapor-phase deposition chamber for DDT exposure, as described previously4. Other samples were transferred directly into the chamber immediately after UV-ozone treatment.

After loading substrates into the chamber, the chamber was purged with N2 for 10 min to remove any O2 or H2O that was introduced during sample loading. The chamber temperature was maintained at 70 °C by warm-wall heating with heating tape elements. A glass vial attached to the chamber was used to store the DDT and a valve attached between the bubbler and the chamber controlled the dosage. The DDT bubbler and the line were both heated at 55 °C to increase the DDT vapor pressure and eliminate condensation within the line. After purging for 10 min, the chamber was pumped down to base pressure (about 2 mTorr) before closing off the pump and dosing DDT for 30s. No observable increase of pressure was 4 ACS Paragon Plus Environment

Page 5 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

observed during the dose time. After dosing, the DDT valve was closed, the pump valve was reopened, and the samples were purged for 3 minutes before being removed for analysis.

Non-contact mode atomic force microscopy (AFM) was performed on a Park System NX-10 with an NSC35 tip purchased from Nanoworld. X-ray photoelectron spectroscopy (XPS) was performed on a PHI Versaprobe 3 with an Al Kα X-ray source. Fine scans were collected in high resolution mode with pass energy of 55 eV, a 200 μm spot size, 50 W, and 15 kV. Three to six scans per region were taken with step sizes of 0.1 eV/step, 20 ms per step. Scanning X-ray Imaging (SXI) was performed on the same equipment with a 9 micron spot size and a beam current of 1.0 W at 15 kV. XPS was then performed on specific locations on the SXI image using a 10 μm spot size, 1.2 W, and a 55 eV pass energy. All spectra were shifted so that the carbon peaks aligned with the aliphatic carbon binding energy of 284.8 eV.

Cross-sectional transmission electron microscopy (TEM), electron energy loss spectroscopy (EELS), and electron dispersive spectroscopy (EDS) were prepared, imaged and analyzed all at Nanolab Technologies, Inc, Milpitas, CA. Samples were first synthesized at Stanford and quickly sealed inside a vacuum transport system before being transferred to Nanlolab Technologies to prevent extended exposure to air. There, TEM lamella were prepared by a FEI Helios Nanolab 450S SEM/Focused Ion Beam (FIB) dual beam system and imaged on a JEOL 2010F TEM at 200kV. The lamella were then transferred to a probecorrected FEI Titan for EELS/EDS analysis at 200kV. The Titan TEM used a ChemiSTEM FEI X-FEG high brightness Schottky field emission gun, a Super-X 4-SDD, windowless EDS detector system, and a Gatan Enfinium ER EELS spectrometer with a fast shutter. Simultaneous EDS/EELS spectrum images were collected to cover all the elements of interests. When performing cross-sectional atomic compositional analyses, EELS was used to detect carbon, silicon, tantalum, copper, and oxygen; EDS was used to detect sulfur.

5 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 32

FTIR spectroscopy measurements were performed using a Nicolet iS50 FTIR spectrometer with a diamond attenuated total reflectance (ATR) plate. Spectra were averaged over 200 scans with 4 cm−1 resolution. An equivalent UV-ozone or acetic acid treated copper-on-silicon substrate was used as a background.

Grazing incidence X-ray diffraction (GIXRD) measurements were performed on beamline 11-3 at the Stanford Synchrotron Radiation Lightsource (SSRL) in a helium-purged chamber with Kapton windows, using a MAR345 imaging plate area detector with X-ray energy of 12.7 keV. The detector distance of around 300 mm and detector orientation were calibrated using a LaB6 standard. Measurements were performed at an incidence angle of 0.12° relative to the X-ray beam. After preparation, samples were transported immediately to the beamline in a nitrogen-filled sealed container.

Scanning electron microscopy (SEM) was performed using an FEI XL30 Sirion SEM with beam energy of 5 kV. Samples were mounted onto a conductive stage using Electrodag 502 purchased from Ted Pella. Inc. Particle size distribution analysis was performed using ImageJ software. Note that the software was generally found to under-size the particles and occasionally misrepresented one particle as two. As a result, the standard deviation in the particle area was high, although total percent surface coverage was generally found to be more consistent.

WCA measurements were performed on an FTA 200 goniometer using deionized (Millipore) water. Auger electron spectroscopy (AES) mapping was performed using a PHI 700 Scanning Auger Nanoprobe with beam settings 10 kV and 10 nA. Maps were obtained using ten scans with a pixel resolution of 128 x 128.

Results and Discussion

6 ACS Paragon Plus Environment

Page 7 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

To examine the effect of DDT exposure onto cupric oxide (CuO), vapor-phase DDT was exposed to Cu/SiO2 patterned substrates and examined with AFM, XPS, and TEM. The use of patterned substrates was necessary in order to observe changes in thickness as a result of the DDT exposure; on these substrates, the silicon region, which has previously been shown not to form SAMs after DDT exposure6, served as a region of fixed height for reference.

Figure 1a shows a representative AFM image of the Cu/SiO2 patterned samples exposed to DDT. The square areas are Cu with DDT deposited on the surface; the other regions are SiO2. From this image, a line scan of the height was taken (Figure 1b, black). Also shown in Figure 1b for comparison is a line scan from an acetic acid etched Cu/SiO2 patterned wafer without any DDT, taken over the same area of the pattern (grey); a representative AFM micrograph of the etched sample is shown in Supplemental Figure 1. Because it is known that vapor-phase thiols can etch and reduce the CuO4,5,11, the acetic acid etched control can be used to determine the baseline elevation of the bare Cu metal underlying the DDT. The line scan from the acetic acid-treated sample shows that the copper regions sit at a depth of about 6 nm from the top surface of the sample. The line scan from the DDT-coated substrate shows that the same square regions have an average height of around 2 nm compared to the silicon regions, indicating that the thickness of the DDT layer is ~8 nm in total, well above the expected molecular length of DDT (~ 2 nm)29. The copper oxide thickness before DDT exposure was roughly 3 nm, based on AFM measurements (not shown).

7 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 32

Figure 1: a) AFM micrograph showing a topographical map of the copper-on-silicon substrates after the vapor deposition of dodecanethiol. b) Line scans of the DDT-coated substrate shown in (a) (black) and from a substrate etched with acetic acid without any DDT (grey). c) XPS fine scan of the UV-ozone treated copper surface before (grey) and after (black) dodecanethiol deposition. d) Cross-sectional SEM images of a region of the patterned substrate in (a). The green box represents the approximate region used in (e). e) Cross-sectional EELS/EDS of a region of the samples after vapor-phase dodecanethiol deposition.

Figure 1c shows a Cu 2p3 XPS fine scan before and after DDT deposition. Before DDT exposure, copper peaks are observed at 932.5 eV, 934.5 eV, along with a satellite peak between 938–946 eV. The peak at 932.5 eV arises from Cu metal beneath the surface oxide layer. The peak at 934.5 eV and the satellite peak between 938 and 946 eV are from CuO, demonstrating that UV-ozone exposure forms Cu2+ oxide5,30,31. However, after subsequent exposure to DDT, the oxide peaks are removed, leaving only a single Cu peak at 932.5 eV associated with Cu0 or Cu1+ (the oxidation state between these cannot be easily distinguished by XPS)11,32. The removal of the CuO features is consistent with previous reports that

8 ACS Paragon Plus Environment

Page 9 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

thiols can reduce and etch the CuO surface4,5,8–15. This result also demonstrates that the DDT is able to penetrate the entirety of the CuO layer and not just the outermost surface.

To better understand the composition of the 8 nm-thick DDT layer, cross-sectional dark field TEM images were taken of the patterned Cu/SiO2 after DDT deposition (Figure 1d). A layer of material ~ 7 nm thick is observed on top of the copper, consistent with the thiol-layer thickness described above. EELS/EDS analysis was performed on a smaller region of the scan (Figure 1e) to probe for C, O, S, Cu, Si, and Ta (here, TaN is used as a liner to reduce copper diffusion into the silicon). The EELS/EDS analysis shows the presence of the Si substrate, the Ta barrier layer, and the Cu square region. Filling most of the void space above the substrate is a carbon-based glue layer, used to protect the surface during the FIB process. Of note is the area directly above the copper region, which shows a sulfur signal nearly 8 nm thick. This region also includes a slight change in C contrast from the background carbon-based glue. These signals are consistent with a film comprised of DDT molecules. Interestingly, a weak copper signal is also present above the copper substrate in this region, suggesting that some copper has been incorporated into the thiolate film.

Combined, the results of both AFM and TEM/EDX suggest that exposure of the CuO surface to DDT leads to 1) the formation of an 8 nm-thick film, thicker than would be expected for a DDT monolayer, 2) the reduction of copper from Cu2+ to either Cu0 or Cu1+, along with the removal of oxygen, and 3) a film composition that includes carbon, sulfur, and copper. These results support the idea of vapor-phase formation of copper-thiolate multilayers, similar to that previously reported for liquid-phase thiol exposure to CuO. Of the two possible models proposed in the literature for this multilayer formation in solution phase [i.e. complexing of the thiols with copper ions from (1) copper ion dissolution into the solvent and re-adsorption,17 or (2) non-dissolved copper ions of the partially reduced copper oxide11], only the second model would be applicable here in the solvent-free, vapor phase process.

9 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 32

The structure of the Cu-thiolate multilayer, including whether or not it is composed of self-assembled regions or is largely amorphous, is not readily obtainable from AFM or TEM. Therefore, GIXRD was performed on DDT-coated blanket copper substrates in order to further elucidate the Cu-thiolate structure. A representative GIXRD pattern is shown in Figure 2a. Interestingly, multiple peaks along both the qxy and qz directions are present, suggesting that the films contain ordering both parallel and perpendicular to the substrate surface. Figures 2b and 2c show integrated regions along the qxy and qz directions respectively. Many of these peaks appear uniformly spaced, with a distance between peaks of 0.177 Å-1 (Supplemental Figure 2). The uniform spacing of these peaks suggests that they correspond to the (0k0) peaks of a lattice with spacing 2𝜋⁄Δ𝑞 = 35.5 Å. In addition, a peak in the qxy direction at 1.64 Å-1 (dspacing of 3.8 Å) is present, which does not correspond to one of the (0k0) peaks and has a higher peak width. Instead, this peak may indicate a less-well-ordered lattice along the (h00) direction. The origin of these peaks is discussed further below.

10 ACS Paragon Plus Environment

Page 11 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 2: a) GIXRD of dodecanethiol-on-copper substrates immediately after deposition. b,c) integrated intensity of the GIXRD patterns along the b) qxy and c) qz directions. The sharp spikes are artifacts of the detector.

11 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 32

Previous work has reported the structure of Cu-thiolate crystallites with different carbon chain lengths33, based on a bulk wet synthesis approach. There, Cu-thiolates liquid crystallites were assigned to a head-tohead bilayer-type structure, where two planes of copper atoms are sandwiched between two monolayers of thiolates. Although the authors do not report a crystal structure for in-plane alkyl chains, it is reasonable to assume that a typical c(2 x 2) structure34 for thiols-on-copper would form. The reported (0k0) distance between bilayers was found to be related to the equation33: 𝑑 = 7.960 + 2.439𝐿, where d is the distance between bilayers and L is the number of CH2 groups in the thiol chain. Given the 11 methylene groups in DDT, the published model predicts a crystallite d-spacing of 34.8 Å, which agrees well with the 35.5 Å spacing measured above.

Based on previous investigations of the copper-thiolate crystal structure33, it is reasonable to infer that the sharp, uniformly spaced peaks represent the well-ordered spacing between laminar layers of the copperthiolate. These lattices are well textured towards the in-plane and out-of-plane directions, with no ordering at angles other than 0° or 90° from the surface normal, since the peaks are all oriented along either the qxy and qz directions. The broader peak at qxy = 1.64 Å-1 (d-spacing of 3.8 Å) likely arises from the chain-chain spacing between alkane groups oriented out-of-plane (typically34 around 4 Å for alkane chains). The expected analogous peak in the qz direction, corresponding to the chain-chain spacing between crystallites lying flat across the substrate surface, may simply be unobservable in GIXRD due to the experimental geometry35.

From the results presented here and previous work on the structure of copper-thiolate crystals33, a representation of the crystal structure formed during the vapor-phase deposition of dodecanethiol on cupric oxide can be constructed, as shown in Figure 3 below. These multilayers take the form of welltextured crystallites likely composed of head-to-head and tail-to-tail thiols groups bound to Cu1+ ions and are oriented either horizontally or vertically relative to the substrate. The Cu ions are inferred to be in the 1+ oxidation state because XPS results above suggest they are not in the 2+ oxidation state and the thiols 12 ACS Paragon Plus Environment

Page 13 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

are expected to be bound at least once to Cu, eliminating the possibility of the Cu0 metal state. The incorporation of these copper ions is expected to occur as a result of the etching of the copper oxide, which releases the copper ions and allows them to be combined with thiols to form the copper thiolate multilayers.

Figure 3) Illustration of the horizontally and vertically aligned multilayer structures of copper-dodecanethiolate.

The mechanistic origin of the two alignments (vertical and horizontal) is not immediately apparent from the data presented here. However, they may be reasonably explained through the consideration of SAM formation in the absence of multilayers. Previous studies have shown that thiols initially nucleating on Cu metals lay flat against the metal surface due to van der Waals forces between the chain and the substrate34,36. After a critical surface concentration is achieved, thiols self-assemble on the metal into vertically aligned chains so that the sulfur atoms may bind to the substrate while the alkyl chains pack tightly into crystalline regions. Here, the crystallites may adopt similar horizontal and vertical orientations, due to the initial adsorption of chains into either horizontal or vertically aligned segments, which then act as nucleation points for the crystallites.

The narrow peak widths along the qxy direction (0.038 Å-1, see Supplemental Figure 2a), suggest a surprisingly well-aligned crystalline structure, with coherence of length 2𝜋⁄𝐹𝑊𝐻𝑀 = 165 Å, spanning many crystal planes. In contrast, the out-of-plane peak widths (0.1 Å-1, see Supplemental Figure 2b) are broader and only show coherence of around 63 Å, likely because the qz direction is limited by the total

13 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 32

film thickness. The broad in-plane chain-chain packing peak at qxy = 1.64 Å-1 has a peak width of 0.15Å-1 and a coherence length of 42 Å, suggesting that either the chain-chain packing is slightly less wellordered than the end-to-end in-plane horizontal crystallites or that the out-of-plane crystallites have a lateral size no longer than around 4 nm.

Previous work has suggested that multilayer formation only occurs on cupric oxide. To examine this for vapor-phase DDT, samples with the native oxide layer removed were prepared by exposing the cleaned samples to acetic acid before DDT deposition. Figure 4a shows a Cu 2p3 XPS high-resolution scan of an acetic acid-etched, blanket copper substrate before and after DDT deposition. The absence of any Cu oxide peaks at 934.5 or 938-946 eV indicates that the oxide layer has been successfully removed by the acetic acid etch. As expected, XPS of the sample after DDT exposure also shows no copper oxide peaks. Patterned Cu/SiO2 substrates were then prepared using a similar acetic acid treatment and examined with AFM, GIXRD, and cross-sectional TEM/EELS/EDS, as above. Figure 4b shows the AFM line scan of the acetic acid-treated Cu/SiO2 patterned substrate before (grey) and after (black) DDT deposition (a topographical micrograph is shown in Supplemental Figure 3). In contrast to the deposition on CuO, DDT onto acetic acid-treated Cu shows an increase in thickness of only about 2 nm, close to the thickness expected for a monolayer of DDT29. GIXRD patterns (Figure 4c) of the blanket copper samples also do not show any of the peaks previously seen for the multilayer samples. Although a well-formed 2D SAM would have substantial in-plane order, qxy peaks associated with that ordering may simply be too low in intensity to detect.

14 ACS Paragon Plus Environment

Page 15 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 4: a) Cu XPS after acetic acid treatment (grey) and after subsequent vapor-phase DDT exposure (black). b) A line scan extracted from AFM micrograph of the acetic acid-treated copper-on-silicon substrates before (grey) and after (black) dodecanethiol vapor deposition. c) GIXRD of acetic acid-treated dodecanethiol-on-copper substrates immediately after deposition. d) FTIR spectra of a copper blanket substrate after DDT exposure with and without a pretreatment with acetic acid. Inset shows a magnified view of the C-H stretching region with the spectra renormalized. e) Cross-sectional TEM, EELS, and EDS spectral images of the patterned substrate in (b) after vapor-phase dodecanethiol deposition.

Cross-sectional TEM and EELS/EDS mapping (Figure 4e) again show the presence of a thin layer of material on top of the copper. However, this layer is much thinner than before, consistent with AFM measurements of the layer thickness. EELS/EDS spectra also show that this layer no longer contains copper and instead shows a thin sulfur region near the surface. Thus, in the case of DDT exposure onto copper metal, the data are consistent with a model in which thiol molecules no longer etch into the

15 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 32

surface, preventing the formation of multilayer structures. As a result, only a monolayer of DDT is formed.

It is interesting to understand what limits the ultimate thickness of the thiol-on-copper oxide multilayer film. Given that the multilayer structure appears to result from the thiol etching of copper oxide, the thickness of the oxide layer is likely to have a strong impact on the thickness of the resulting multilayer film. The laminar structure of the copper-thiolate crystallites further suggests that the thiol aliphatic tail length is likely to affect the multilayer thickness. We can roughly estimate the expected thickness of a Cu-thiolate film formed from a copper oxide layer of a given thickness if we assume that the copper atoms in the thiolate films come entirely from those contained in the copper oxide. By assuming a C-C spacing of ≈3.8–4 Å between the thiol aliphatic tail groups and a repeat bilayer thickness of 35.5 Å, we can estimate a density for the Cu-thiolates of ≈4–5 copper atoms per nm3. In comparison, the copper density in copper oxide is ~ 48 copper atoms per nm3, based on a bulk density of 6.31 g/cm3. Hence, a 3 nm thick CuO film would be expected to give rise to an upper limit of 29 nm for the Cu-thiolate film thickness, if there were no loss of Cu atoms from the film. This value is within a factor of 4 of the measured film thickness of 8 nm, providing further justification that the oxide thickness plays a direct role on the resulting Cu-thiolate film thickness. The lower copper density in the Cuthiolate could be the result of Cu lost during the etching process as a volatile product. Ultimately, additional experiments are needed to fully understand how these properties affect the resulting multilayer thickness. Such experiments could include modulating the copper oxide thickness and the chain length of the exposed thiols.

The SAM quality and degree of packing within SAMs is often measured with techniques like WCA and FTIR spectroscopy4,5. For example, as the SAM reaches a dense layer with tight packing, the WCA approaches a plateau, which has been reported to occur at a value near 110° for DDT on copper5. FTIR spectroscopy is also sensitive to the crystallinity of the packing of the alkane tails of thiols. For wellformed, highly crystalline SAMs, the ideal9,29,37 symmetric (νsym(CH2)) and asymmetric (νasy(CH2)) CH2

16 ACS Paragon Plus Environment

Page 17 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

stretching modes are ~ 2850 cm-1 and 2917 cm-1. These two peaks shift to higher wavenumbers if the SAM is less ordered and more liquid-like29.

To compare against the properties of well-formed SAMs, the properties of the DDT layers on copper with and without acetic acid pre-treatment were also examined using WCA measurements and FTIR spectroscopy. Despite their major differences in thickness and composition, both multilayer and monolayer samples showed the same water contact angle of 106 ± 1° (n = 4, where n is the number of separate samples measured). However, the roughness of the copper-thiolate multilayer film (1.2 nm rms, Supplemental Figure 4c) is much higher than the roughness of the thiol-on-copper monolayer5 (0.5 nm rms, Supplemental Figure 4f). Increased roughness has been reported to make hydrophilic and hydrophobic surfaces exhibit lower and higher WCAs respectively38, which may contribute to a higher WCA for the copper-thiolate multilayer film than would be expected from just the chemistry. FTIR spectra of the CH2 stretching region for the samples are shown in Figure 4d. Both spectra show the expected symmetric and antisymmetric CH2 stretching peaks (2853 cm-1 and 2925 cm-1 for monolayer DDT; 2851 cm-1 and 2920 cm-1 for multilayer Cu-thiolate, respectively), as well as peaks associated with the terminal CH3 group’s symmetric (νsym(CH3)) and antisymmetric (νasy(CH3)) stretching (2876 cm-1 and 2965 cm-1 for monolayer DDT; 2873 cm-1 and 2960 cm-1 for multilayer Cu-thiolate, respectively). The CH3 group stretching frequencies agree well with literature values of 2878 cm-1 and 2965 cm-1, respectively15. The intensity of the CH2 stretching region for the multilayer film is much higher than for the monolayer, in agreement with the expected increase in the number of CH2 groups in the multilayer film. Interestingly, the peaks of the multilayer structure are slightly less blue-shifted than the monolayer structure, indicating that the multilayer structure is "more crystalline" than the monolayer. The CH3/CH2 peak intensity ratio for the monolayer also differs noticeably from the 1:11 ratio expected for DDT. This has been observed before for SAMs on metal and is explained as a result of the dipole moment cancellation from the metal surface, which attenuates the CH2 modes15. The Cu-thiolate film would be expected to experience less of this effect, as the film both is thicker and has components in both 17 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 32

horizontal and vertical orientations. These data on both monolayer and multilayer films suggest that, though both films appear to be reasonably well-ordered, care must be taken in confirming the layer thickness in the Cu-thiolate system using WCA and FTIR spectroscopy, since neither method appears to be sensitive to the film thickness. Other complementary methods must be carried out to distinguish between monolayer and multilayer films.

Because several key applications of thiolate SAMs on Cu require that they serve as effective passivation layers to protect or block the underlying copper23,25–27, the stability of the layers against various chemical agents including ambient atmosphere is important. Of particular interest is the stability of the Cu-thiolate multilayers compared to the monolayer. Here, we investigated the stability of the Cu-thiolate multilayers in air. DDT-coated copper blanket substrates were left in a) ambient air and b) a nitrogen purged storage container for several days and examined with SEM, Auger electron spectroscopy, GIXRD, and XPS. Samples stored in the nitrogen purging environment for two weeks showed negligible changes in appearance and no distinct features in SEM scans. In contrast, samples stored in air undergo large-scale coarsening into micron sized particles of Cu-thiolate, as described below. These particles could also be seen by eye. Monolayer DDT films deposited onto acetic acid-treated copper blanket substrates did not show the presence of any particles by SEM after several weeks in air.

SEM images taken of multilayer samples exposed to air for as little as one day (Figure 5a) developed dark patches as large as 20 µm2 corresponding to particles. Near the edge of 1 cm2 copper samples, longer dendritic particles were also visible (Figure 5b). Higher magnification of a particle is shown in Figure 5c. In this image, the edges of the particle are shown to be faceted, with several edges appearing to intersect one another at 120° angles, consistent with possible crystallite formation. AFM images taken of the particles (Figure 5d, Supplemental Figure 5), show that they have a thickness of a few hundred nm. This further corroborates the formation of particles rather than vacancy islands, as has been seen for alkanethiols deposited on gold39–41. 18 ACS Paragon Plus Environment

Page 19 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

The elemental composition of these particles after 4 days in air was examined using SXI/XPS. Scans collected on top of a particle show the same composition as scans taken in regions without a particle (Figure 5e,f). In both cases, the Cu 2p fine scans show a single peak around 932.3 eV, consistent with copper metal, and the appearance of a slight shoulder around 935 eV. The S 2p fine scans shows two sets of peaks: thiol sulfur at 162.1 eV (2p3/2) and 163.2 eV (2p1/2), as well as a higher binding energy peak at 168 eV. The principal peaks therefore arise from the copper substrate and the DDT molecules, indicating that the particles as well as the film around it is largely composed of copper thiolate. However, the combination of the Cu 2p peak at 935 eV and the S 2p peak at 168 eV indicates that there is also some contribution from copper sulfonate, CuSO2C12H26, as has been assigned in the literature25,42,43. Two weeks after crystallite formation, GIXRD also shows no major changes in the (0k0) peaks associated with the head-to-head structure, though the chain-chain stacking peak has largely disappeared (Supplemental Figure 6). This suggests that the conversion of the thiols to sulfonates does not disrupt the layered structure shown previously (Figure 3) but may disrupt the packing between neighboring carbon chains.

19 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 32

Figure 5: a,b) SEM images of the thiol multilayer films after exposure to air for two weeks. c) A higher magnification image of a thiol crystallite. Dashed guidelines are added to highlight crystallite facets. d) AFM image of a crystallite, showing it has a height around 200 nm. e, f) SXI/XPS scans taken of a crystallite (red) and a non-crystallite (blue) region of a sample.

To better understand the process by which the thiolate multilayer transforms into particles of copper thiolate, SEM images of three copper oxide films exposed to vapor DDT were taken 0, 1, 2, 4, 8, and 16 days after exposure to air. Representative images of the films are shown in Supplemental Figure 7. Films immediately taken to the SEM after DDT exposure show no indication of particle formation. However, after a day in air, particles ~ 5 μm in diameter become apparent. Based on these observations, a particle size analysis was performed using the ImageJ software44,45. The resulting average particle areas and percent particle surface coverage are shown in Figure 6a and 6b respectively. The data show that after the first day, the particles continue to grow in size until reaching a diameter of around 7 μm (day 4);

20 ACS Paragon Plus Environment

Page 21 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

additional air exposure beyond day 4 does not result in a significant increase in particle size or percent surface coverage.

The data suggest that copper thiolate multilayer is undergoing ripening, in which copper thiolate groups dewet from bulk regions and aggregate together into larger islands to reduce defects and minimize surface energy, as has been seen for many kinds of thin films46. An application of the particle area data to a simple time-dependent model (𝑟 = 𝐵𝑡 1⁄𝑛 ) suggests that the particle sizes grow with a time constant of n = 5.2 ± 2.8 (fit is shown as the dashed line in Figure 6b). Although this is only a rough model, the time constant is within the range expected for the 2D diffusion of molecules into hemispherical crystallites47,48 (r ~ t1/4) and the time constant expected for the dewetting of a film from a straight front (r ~ t1/5)49,50. We postulate that the slowing of continued particle growth after four days results from depletion of the multilayer: as the particles continue to nucleate, the thickness of the multilayer decreases as molecules from the layers diffuse into the aggregates, eventually leading to the consumption of the film and a slowing of the particle growth. A numerical estimate of the film consumption (see supporting information) suggests that the particle growth will slow by ~4 days because a significant portion of the multilayer chains will have already aggregated into particles, consistent with the experimental time constant. However, experimental confirmation of the loss of multilayer thickness after particle ripening was not performed, since a substrate like the one shown in Figure 1b for AFM measurement, which was necessary to measure the film height relative to a fixed position, could not be obtained with a large enough copper region to fit the entirety of a particle (~20 microns long).

21 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 32

Figure 6) Plots of a) the percent coverage and b) the average area of the particles on the copper surface after exposure to air as a function of time. The dashed curve represents a fit of the particle area to a model (𝑟 = 𝐵𝑡 1⁄𝑛 ), resulting in a time constant of 𝑛 = 5.2. The model only includes data through day 4 because the ripening of the film terminates once the film is fully consumed.

XPS analysis was also performed as a function of air exposure time for the copper oxide blanket substrates exposed to DDT. The copper and sulfur fine scans are shown in Figures 7a,b. Copper 2p fine 22 ACS Paragon Plus Environment

Page 23 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

scans show a single peak at 932.3 eV through day 2, indicating that the exposed copper remains primarily in the Cu1+ oxidation state during this time and is not oxidized. By day 4, a shoulder peak consistent with sulfone around 935 eV, as well as a higher binding energy peak around 943 eV (Cu2+) become visible, suggesting some degree of oxidation has occurred. In the S 2p fine scan, only the thiol 2p3/2 and 2p1/2 peaks at 162.1 eV and 163.2 eV, respectively, are visible initially. Again, by day 4, an additional sulfone peak around 168 eV becomes apparent. Thus, as the films are exposed to air, they slowly begin to degrade and become oxidized, converting the thiol to sulfone and Cu1+ to Cu2+. This rate of oxidation is comparable to what has been seen previously25 as well as measured here (Supplemental Figure 8) for monolayer thiols on copper. However, because the oxidation state of the underlying copper is difficult to distinguish from the oxidation state of the copper within the multilayers, it is unclear if the multilayer is able to prevent surface oxidation or not. At the very least, the multilayer is no better at preventing oxidation of the copper within the multilayers than the copper underneath a DDT monolayer.

However, while exposure to air seems necessary to induce particle ripening (no particles form on samples stored in a nitrogen purging environment), the particle ripening is not driven purely by the oxidation of the multilayer, as evidenced by the following two observations: 1) the timescale of particle ripening does not coincide with the timescale of the copper oxidation, as considerable particle ripening has taken place by day 4 when very little copper or sulfur oxidation has taken place, and 2) the particles do not show a greater degree of oxidation than the surrounding multilayer (Figure 5e, f). Further studies are thus needed to investigate the ultimate driving force toward particle formation.

23 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 32

Figure 7) XPS fine scans of the a) copper 2p and b) sulfur 2s peaks of the CuO substrates exposed to DDT as a function of air exposure time between 0 and 16 days.

24 ACS Paragon Plus Environment

Page 25 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Conclusions In this work, we demonstrate that the vapor-phase deposition of dodecanethiol onto copper oxide results in the formation of several nanometer thick Cu-thiolate multilayer films, rather than self-assembled monolayers. These multilayers only form when a surface CuO is present, verifying that the etching of the copper oxide films by the thiol molecules is a key step in the multilayer formation. Given the lack of solvent present in the vapor-phase system, this further suggests that the mechanism of multilayer formation does not revolve around the dissolution and renucleation of copper ions, as has been proposed previously for solution processes17, but involves the coordination between undissolved Cu1+ ions and thiol molecules11. Pre-removal of the oxide with acetic acid, in contrast, prevents the formation of multilayers and instead leads to self-assembled monolayers.

Multilayer films are found to be crystalline, with a lattice spacing of 35.5 Å, consistent with a bilayer thiol structure sandwiched between two layers of copper atoms. The crystallites are also well-ordered, with domains lying flat against the underlying copper surface and domains standing vertically out of the plane of the substrate.

Exposure of the multilayer films to air results in a ripening of the copper-thiolate crystallites into micronsized particles. The formation of these particles, which have similar composition to the multilayers, implies that the multilayer films have sufficient chain mobility to restructure into particles. This ripening forms on the same timescale as seen for ripening from 2D to 3D structures and is believed to stop once the multilayer film becomes depleted from the regions around a particle. The copper and thiol within these particles also appears to oxidize after air exposure, though on a different timescale than the particle formation. However, this oxidation occurs at the same rate as the oxidation seen for copper passivation with monolayers, suggesting that the multilayer provides no additional benefit to surface passivation than ordinary monolayer films, other than serving as a sacrificial oxidation layer. Combined, these results provide new fundamental insight into the vapor-phase formation of copper thiolate multilayers, the 25 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 32

molecular level structure of those films, and the mechanism behind their ripening into particles during air exposure.

Acknowledgements This work was supported by the National Science Foundation (Grants CHE-1607339 and CHE-1213879). D.S.B. would also like to acknowledge support from an NSF Graduate Research Fellowship and the Gerald J. Lieberman Fellowship provided by the Stanford Vice Provost for Graduate Education. Part of this work was performed at the Stanford Nano Shared Facilities (SNSF), supported by the National Science Foundation under award ECCS-1542152. Use of the Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory, is supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract No. DE-AC02-76SF00515.

Supporting Information AFM micrographs of patterned Cu/SiO2 surfaces and blanket copper substrates exposed to various conditions, plot of peak positions and peak widths from GIXRD patterns, AFM line scan and GIXRD pattern of particle resulting from dewetting, SEM images of multilayers after air exposure showing particle formation, XPS spectra of acetic-acid treated CuO substrates after air exposure, numerical estimation of film thickness after film dewetting.

26 ACS Paragon Plus Environment

Page 27 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

References (1)

Vericat, C.; Vela, M. E.; Corthey, G.; Pensa, E.; Cortés, E.; Fonticelli, M. H.; Ibañez, F.; Benitez, G. E.; Carro, P.; Salvarezza, R. C. Self-assembled monolayers of thiolates on metals: a review article on sulfur-metal chemistry and surface structures. RSC Adv. 2014, 4, 27730–27754 DOI: 10.1039/C4RA04659E.

(2)

Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M. Self-assembled monolayers of thiolates on metals as a form of nanotechnology; 2005; Vol. 105.

(3)

Ulman, A. Formation and structure of self-assembled monolayers. Chem. Rev. 1996, 96, 1533– 1554 DOI: 10.1021/cr9502357.

(4)

Hashemi, F. S. M.; Bent, S. F. Sequential Regeneration of Self-Assembled Monolayers for Highly Selective Atomic Layer Deposition. Adv. Mater. Interfaces 2016, 3, 1600464 DOI: 10.1002/admi.201600464.

(5)

Minaye Hashemi, F. S.; Birchansky, B. R.; Bent, S. F. Selective Deposition of Dielectrics: Limits and Advantages of Alkanethiol Blocking Agents on Metal-Dielectric Patterns. ACS Appl. Mater. Interfaces 2016, 8, 33264–33272 DOI: 10.1021/acsami.6b09960.

(6)

Prasittichai, C.; Pickrahn, K. L.; Minaye Hashemi, F. S.; Bergsman, D. S.; Bent, S. F. Improving Area-Selective Molecular Layer Deposition by Selective SAM Removal. ACS Appl. Mater. Interfaces 2014, 6, 17831–17836 DOI: 10.1021/am504441e.

(7)

Mackus, A. J. M.; Bol, A. A.; Kessels, W. M. M. The use of atomic layer deposition in advanced nanopatterning. Nanoscale 2014, 6, 10941–10960 DOI: 10.1039/C4NR01954G.

(8)

Carbonell, L.; Whelan, C. M.; Kinsella, M.; Maex, K. A thermal stability study of alkane and aromatic thiolate self-assembled monolayers on copper surfaces. Superlattices Microstruct. 2004, 36, 149–160 DOI: 10.1016/j.spmi.2004.08.033.

(9)

Dilimon, V. S.; Denayer, J.; Delhalle, J.; Mekhalif, Z. Electrochemical and Spectroscopic Study of the Self-Assembling Mechanism of Normal and Chelating Alkanethiols on Copper. Langmuir 2012, 28, 6857–6865 DOI: 10.1021/la300021g. 27 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(10)

Page 28 of 32

Hosseinpour, S.; Gothelid, M.; Leygraf, C.; Johnson, C. M. Self-Assembled Monolayers as Inhibitors for the Atmospheric Corrosion of Copper Induced by Formic Acid: A Comparison between Hexanethiol and Hexaneselenol. J. Electrochem. Soc. 2013, 161, C50–C56 DOI: 10.1149/2.056401jes.

(11)

Wang, Y.; Im, J.; Soares, J. W.; Steeves, D. M.; Whitten, J. E. Thiol Adsorption on and Reduction of Copper Oxide Particles and Surfaces. Langmuir 2016, 32, 3848–3857 DOI: 10.1021/acs.langmuir.6b00651.

(12)

Calderón, C. A.; Ojeda, C.; Macagno, V. A.; Paredes-Olivera, P.; Patrito, E. M. Interaction of Oxidized Copper Surfaces with Alkanethiols in Organic and Aqueous Solvents. The Mechanism of Cu 2 O Reduction. J. Phys. Chem. C 2010, 114, 3945–3957 DOI: 10.1021/jp9045148.

(13)

Azzaroni, O.; Cipollone, M.; Vela, M. E.; Salvarezza, R. C. Protective Properties of Dodecanethiol Layers on Copper Surfaces: The Effect of Chloride Anions in Aqueous Environments. Langmuir 2001, 17, 1483–1487 DOI: 10.1021/la000852c.

(14)

Sung, M. M.; Sung, K.; Kim, C. G.; Lee, S. S.; Kim, Y. Self-Assembled Monolayers of Alkanethiols on Oxidized Copper Surfaces. J. Phys. Chem. B 2000, 104, 2273–2277 DOI: 10.1021/jp992995h.

(15)

Ron, H.; Cohen, H.; Matlis, S.; Rappaport, M.; Rubinstein, I. Self-Assembled Monolayers on Oxidized Metals. 4. Superior n -Alkanethiol Monolayers on Copper. J. Phys. Chem. B 1998, 102, 9861–9869 DOI: 10.1021/jp9827072.

(16)

Lecordier, L.; Herregods, S.; Armini, S. Vapor-deposited octadecanethiol masking layer on copper to enable area selective Hf3N4 atomic layer deposition on dielectrics studied by in situ spectroscopic ellipsometry. J. Vac. Sci. Technol. A 2018, 36, 031605 DOI: 10.1116/1.5025688.

(17)

Keller, H.; Simak, P.; Schrepp, W.; Dembowski, J. Surface chemistry of thiols on copper: an efficient way of producing multilayers. Thin Solid Films 1994, 244, 799–805 DOI: 10.1016/00406090(94)90574-6.

(18)

Schlenoff, J. B.; Li, M.; Ly, H. Stability and Self-Exchange in Alkanethiol Monolayers. J. Am. 28 ACS Paragon Plus Environment

Page 29 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Chem. Soc. 1995, 117, 12528–12536 DOI: 10.1021/ja00155a016. (19)

Laibinis, P. E.; Whitesides, G. M.; Allara, D. L.; Tao, Y. T.; Parikh, A. N.; Nuzzo, R. G. Comparison of the structures and wetting properties of self-assembled monolayers of nalkanethiols on the coinage metal surfaces, copper, silver, and gold. J. Am. Chem. Soc. 1991, 113, 7152–7167 DOI: 10.1021/ja00019a011.

(20)

Chavez, K. L.; Hess, D. W. A Novel Method of Etching Copper Oxide Using Acetic Acid. J. Electrochem. Soc. 2001, 148, G640 DOI: 10.1149/1.1409400.

(21)

Closser, R. G.; Bergsman, D. S.; Ruelas, L.; Hashemi, F. S. M.; Bent, S. F. Correcting defects in area selective molecular layer deposition. J. Vac. Sci. Technol. A Vacuum, Surfaces, Film. 2017, 35, 031509 DOI: 10.1116/1.4980049.

(22)

Walker, P.; Tarn, W. H. CRC Handbook of Metal Etchants; Tarn, W., Walker, P., Eds.; CRC Press, 1990.

(23)

Scherer, J.; Vogt, M. R.; Magnussen, O. M.; Behm, R. J. Corrosion of Alkanethiol-Covered Cu(100) Surfaces in Hydrochloric Acid Solution Studied by in-Situ Scanning Tunneling Microscopy. Langmuir 1997, 13, 7045–7051 DOI: 10.1021/la970372u.

(24)

Jiang, X.; Bent, S. F. Area-Selective ALD with Soft Lithographic Methods: Using Self-Assembled Monolayers to Direct Film Deposition. J. Phys. Chem. C 2009, 113, 17613–17625 DOI: 10.1021/jp905317n.

(25)

Laibinis, P. E.; Whitesides, G. M. Self-assembled monolayers of n-alkanethiolates on copper are barrier films that protect the metal against oxidation by air. J. Am. Chem. Soc. 1992, 114, 9022– 9028 DOI: 10.1021/ja00049a038.

(26)

Hutt, D. A.; Liu, C. Oxidation protection of copper surfaces using self-assembled monolayers of octadecanethiol. Appl. Surf. Sci. 2005, 252, 400–411 DOI: 10.1016/j.apsusc.2005.01.019.

(27)

Feng, Y. Corrosion Protection of Copper by a Self-Assembled Monolayer of Alkanethiol. J. Electrochem. Soc. 1997, 144, 55 DOI: 10.1149/1.1837365.

(28)

Färm, E.; Vehkamäki, M.; Ritala, M.; Leskelä, M. Passivation of copper surfaces for selective-area 29 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 32

ALD using a thiol self-assembled monolayer. Semicond. Sci. Technol. 2012, 27, 074004 DOI: 10.1088/0268-1242/27/7/074004. (29)

Porter, M. D.; Bright, T. B.; Allara, D. L.; Chidsey, C. E. D. Spontaneously organized molecular assemblies. 4. Structural characterization of n-alkyl thiol monolayers on gold by optical ellipsometry, infrared spectroscopy, and electrochemistry. J. Am. Chem. Soc. 1987, 109, 3559– 3568 DOI: 10.1021/ja00246a011.

(30)

Parmigiani, F.; Pacchioni, G.; Illas, F.; Bagus, P. S. Studies of the CuO bond in cupric oxide by Xray photoelectron spectroscopy and ab initio electronic structure models. J. Electron Spectros. Relat. Phenomena 1992, 59, 255–269 DOI: 10.1016/0368-2048(92)87005-7.

(31)

Schön, G. ESCA studies of Cu, Cu2O and CuO. Surf. Sci. 1973, 35, 96–108 DOI: 10.1016/00396028(73)90206-9.

(32)

Mcintyre, N. S.; Cook, M. G. X-Ray Photoelectron Studies on Some Oxides and Hydroxides of Cobalt, Nickel, and Copper. Anal. Chem. 1975, 47, 2208–2213 DOI: 10.1021/ac60363a034.

(33)

Sandhyarani, N.; Pradeep, T. An investigation of the structure and properties of layered copper thiolates. J. Mater. Chem. 2001, 11, 1294–1299 DOI: 10.1039/b009837j.

(34)

Loepp, G.; Vollmer, S.; Witte, G.; Wöll, C. Adsorption of heptanethiol on Cu(110). Langmuir 1999, 15, 3767–3772 DOI: 10.1021/la981510y.

(35)

Baker, J. L.; Jimison, L. H.; Mannsfeld, S.; Volkman, S.; Yin, S.; Subramanian, V.; Salleo, A.; Alivisatos, a P.; Toney, M. F. Quantification of thin film crystallographic orientation using X-ray diffraction with an area detector. Langmuir 2010, 26, 9146–9151 DOI: 10.1021/la904840q.

(36)

Schreiber, F. Structure and growth of self-assembling monolayers. Prog. Surf. Sci. 2000, 65, 151– 256 DOI: 10.1016/S0079-6816(00)00024-1.

(37)

Hong, J.; Porter, D. W.; Sreenivasan, R.; McIntyre, P. C.; Bent, S. F. ALD Resist Formed by Vapor-Deposited Self-Assembled Monolayers. Langmuir 2007, 23, 1160–1165 DOI: 10.1021/la0606401.

(38)

Berg, J. C. An Introduction to Interfaces and Colloids: The Bridge to Nanoscience; WORLD 30 ACS Paragon Plus Environment

Page 31 of 32 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

SCIENTIFIC, 2009. (39)

Cavalleri, O.; Hirstein, A.; Kern, K. Ostwald ripening of vacancy islands at thiol covered Au(111). Surf. Sci. 1995, 340, 960–964 DOI: 10.1016/0039-6028(95)00790-3.

(40)

Edinger, K.; Gölzhäuser, A.; Demota, K.; Wöll, C. H.; Grunze, M. Formation of Self-Assembled Monolayers of n-Alkanethiols on Gold: A Scanning Tunneling Microscopy Study on the Modification of Substrate Morphology. Langmuir 1993, 9, 4–8 DOI: 10.1021/la00025a002.

(41)

Wu, H.; Sotthewes, K.; Kumar, A.; Vancso, G. J.; Schön, P. M.; Zandvliet, H. J. W. Dynamics of decanethiol self-assembled monolayers on Au(111) studied by time-resolved scanning tunneling microscopy. Langmuir 2013, 29, 2250–2257 DOI: 10.1021/la304902y.

(42)

Klein, J. C.; Li, C. P.; Hercules, D. M.; Black, J. F. Decomposition of Copper Compounds in XRay Photoelectron Spectrometers. Appl. Spectrosc. 1984, 38, 729–734 DOI: 10.1366/0003702844555016.

(43)

Louette, P.; Bodino, F.; Pireaux, J.-J. Poly(sulfone resin) XPS Reference Core Level and Energy Loss Spectra. Surf. Sci. Spectra 2005, 12, 100–105 DOI: 10.1116/11.20050921.

(44)

Schindelin, J.; Rueden, C. T.; Hiner, M. C.; Eliceiri, K. W. The ImageJ ecosystem: An open platform for biomedical image analysis. Mol. Reprod. Dev. 2015, 82, 518–529 DOI: 10.1002/mrd.22489.

(45)

Schneider, C. A.; Rasband, W. S.; Eliceiri, K. W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 2012, 9, 671–675 DOI: 10.1038/nmeth.2089.

(46)

Thompson, C. V. Solid-State Dewetting of Thin Films. Annu. Rev. Mater. Res. 2012, 42, 399–434 DOI: 10.1146/annurev-matsci-070511-155048.

(47)

Vengrenovych, R. D.; Ivanskyy, B. V.; Panko, I. I.; Stasyk, M. O.; Fesiv, I. V. Stability of Nanocrystals in 2D and 3D Systems in Ostwald Ripening. Powder Metall. Met. Ceram. 2015, 54, 281–291 DOI: 10.1007/s11106-015-9711-x.

(48)

Shorlin, K.; Krylov, S.; Zinke-Allmang, M. Fundamental problems concerning three dimensional clustering on surfaces. Phys. A Stat. Mech. its Appl. 1998, 261, 248–265 DOI: 10.1016/S037831 ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 32

4371(98)00379-3. (49)

Wong, H.; Voorhees, P. W.; Miksis, M. J.; Davis, S. H. Periodic mass shedding of a retracting solid film step. Acta Mater. 2000, 48, 1719–1728 DOI: 10.1016/S1359-6454(00)00016-1.

(50)

Chame, A.; Pierre-Louis, O. Modeling dewetting of ultra-thin solid films. Comptes Rendus Phys. 2013, 14, 553–563 DOI: 10.1016/j.crhy.2013.06.004.

TOC Figure

32 ACS Paragon Plus Environment