Oxo-Fluoroalkoxide Complexes with Partially ... - ACS Publications

Jul 17, 2019 - deposition temperature and precursor choice.10−13. The properties of these materials depend on the deposition technique and condition...
0 downloads 0 Views 3MB Size
Subscriber access provided by BUFFALO STATE

Functional Nanostructured Materials (including low-D carbon) x

Growth of WO from Tungsten (VI) Oxo-Fluoroalkoxide Complexes with Partially Fluorinated #-diketonate/#-ketoesterate Ligands: Comparison of Chemical Vapor Deposition to Aerosol-Assisted CVD Nathan C Ou, Duane C Bock, Xiaoming Su, Doina Craciun, Valentin Craciun, and Lisa McElwee-White ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.9b08830 • Publication Date (Web): 17 Jul 2019 Downloaded from pubs.acs.org on July 18, 2019

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 59 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

ACS Applied Materials & Interfaces

Growth of WOx from Tungsten (VI) OxoFluoroalkoxide Complexes with Partially Fluorinated β-diketonate/β-ketoesterate Ligands: Comparison of Chemical Vapor Deposition to Aerosol-Assisted CVD. Nathan C. Ou,a Duane C. Bock,a Xiaoming Su,a Doina Craciun,b Valentin Craciunb,c and Lisa McElwee-Whitea,*

aDepartment

bNational

of Chemistry, University of Florida, Gainesville, Florida, 32611-7299 USA.

Institute for Laser, Plasma, and Radiation Physics, Bucharest-Magurele,

Romania

cExtreme

Light Infrastructure for Nuclear Physics, Bucharest-Magurele, Romania

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces 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

Abstract.

Page 2 of 59

Tungsten (VI) oxo complexes of the type WO(OR)3L [R = C(CH3)2CF3,

C(CF3)2CH3,

CH(CF3)2,

L

=

hexafluoroacetylacetonate

(hfac),

ethyl

trifluoroacetoacetonate (etfac), acetylacetonate (acac)] bearing partially fluorinated alkoxide and/or chelating ligands were synthesized. The thermal decomposition behavior and mass spectrometry fragmentation patterns of selected examples were studied. Thermolysis products of WO(OC(CF3)2CH3)3(hfac) were characterized by NMR and GCMS. Studies of the sublimation behavior of the complexes demonstrated that the volatility of the complex depends on the degree of fluorination.

Comparative studies of the

deposition of tungsten oxide by chemical vapor deposition (CVD) and aerosol-assisted (AA)CVD were carried out using WO(OC(CF3)2CH3)3(hfac) as a single source precursor. WOx materials were successfully deposited by both deposition methods, but the deposits differed in morphology, structure and crystallinity.

Keywords chemical vapor deposition, tungsten oxide, nanostructures, volatility, thermolysis

Introduction

ACS Paragon Plus Environment

2

Page 3 of 59 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

ACS Applied Materials & Interfaces

Tungsten oxides (WOx) are well known for their electrochromic properties which have found widespread applications in optoelectronic devices and “smart-window” technology.1,2

Tungsten oxide has also been used in other applications such as

photovoltaics,3 catalysis4 and gas sensors.5

The combination of numerous

stoichiometries and morphologies broadens the range of applications of the material. Tungsten oxide materials can exist in substoichiometric compositions such as WO2.72,6 WO2.66,7 and WO2.8,8 as well as the stoichiometric WO3.9 Different morphologies of WOx, such as nanotubes, nanobelts and nanorods have been obtained by varying deposition temperature, and precursor choice.10-13 The properties of these materials depend on the deposition technique and conditions.14,15 Chemical vapor deposition (CVD) is an attractive deposition technique due to its ability to cover large growth areas with high film conformality.16 Previous reports of CVD of WOx have demonstrated control of composition and morphology to produce both amorphous thin films and nanostructures.17,18 An important step in the CVD process is the volatilization of the precursor, which in traditional CVD involves evaporation of a liquid precursor or sublimation of a solid in the presence of a carrier gas. Therefore, this

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces 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 59

technique requires precursors with significant vapor pressure that are also sufficiently stable to survive the volatilization process. However, the volatility requirement can be circumvented through a variant known as Aerosol-Assisted (AA)CVD. In contrast to traditional CVD, AACVD volatilizes the precursor by nebulization of a solution, so that the carrier gas transports an aerosol of the precursor and solvent to the substrate.19 Once the solvent evaporates, the growth mechanism is understood to be the same as CVD.19 Therefore, the precursor choice becomes dependent on solubility instead of volatility, permitting a wider scope of precursor chemistry to be explored.20 Two precursors commonly used in CVD of tungsten oxide are the homoleptic complexes WCl621 and W(CO)6.22 Although these precursors demonstrate good transport properties, the CVD process produces undesirable, hazardous byproducts such as HCl and CO. They also require a coreactant for oxygen incorporation, making prevention of premature reaction between the coreactants and control of the stoichiometry in the film more challenging.23 As an alternative, the use of single-source precursors facilitates control of film stoichiometry and precursor decomposition through consideration of ligand choice.24 Single-source precursors for WOx CVD have included tungsten alkoxides, such

ACS Paragon Plus Environment

4

Page 5 of 59 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

ACS Applied Materials & Interfaces

as W(OEt)6, but their manipulation is difficult due to their sensitivity to moisture.25,26 This sensitivity can be lessened through use of chelating ligands. Preparation of WOx material through AACVD has been successfully demonstrated by using precursors that contain both alkoxide and chelating ligands.27,28 Fluorination of the ligands can improve the volatility of these complexes due to the repulsive nature of the fluorine atoms.29 By employing this strategy in precursor design, we previously reported the preparation of a series of oxo alkoxide complexes WO(OR)3L where L is a chelating β-diketonate or βketoesterate ligand and the alkoxide ligands -OR are partially fluorinated.30 In this work, we explore the properties of WO(OR)3L complexes where both the alkoxide ligands and the chelating ligand are partially fluorinated. Extending the fluorination to the chelating ligand resulted in a significant increase in precursor stability and volatility with respect to complexes where the fluorination was only on the alkoxide. The increased volatility was sufficient in these complexes for direct volatilization in a conventional CVD process. It is generally understood that AACVD and CVD should involve the same mechanistic steps (after solvent evaporation from the aerosol of the solution)19 and give the same material. Experimental comparisons of material grown by AACVD and CVD

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces 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 59

from the same precursor remain rare. There are isolated examples, such as deposition of WOx materials from W(OEt)5 in the presence of O2 by AACVD31 and atmospheric pressure CVD17 but the reactors and deposition conditions are so different that it is difficult to make a comparison of the two methods. There is a report of growth of fluorine-doped tin dioxide films deposited using dual-source precursors, monobutyl tin trichloride and trifluoroacetic acid, by AACVD and atmospheric pressure CVD (APCVD).32

These

materials grown with different volatilization techniques in similar reactors showed differences in substrate coverage and morphology. However, to our knowledge, a similar comparison of growth of a metal oxide from the same single-source precursor under CVD and AACVD conditions has not been reported. We now report deposition of WOx using WO(OC(CF3)2CH3)3(hfac) as a precursor under both CVD and AACVD conditions. Differences in film composition, morphology and structure were evident, which suggests that solvent plays a critical role during the deposition process.

Experimental Section

ACS Paragon Plus Environment

6

Page 7 of 59 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

ACS Applied Materials & Interfaces

Representative Synthesis:

WO(OC(CF3)2CH3)3(hfac) (5).

In a Schlenk flask,

WO(OC(CF3)2CH3)4 (0.693 mmol, 0.491 g) was dissolved in diethyl ether (40 mL) and chilled in an ice bath for 30 min. Hhfac (0.843 mmol, 0.175 g), dissolved in diethyl ether (10 mL), was added dropwise to the WO(OC(CF3)2CH3)4 solution. Then the reaction was warmed to room temperature and allowed to stir for 15 h. Volatiles were removed in

vacuo to afford a yellowish solid. The solid was sublimed under vacuum (30 mTorr) at 36 °C for 15 h to yield a solid yellowish-white sublimate. Yield: 0.45 g (68 %). 1H NMR (300 MHz, CDCl3, 25 °C): δ 6.49 (s, 1H, OC(CF3)CH), 1.88 (s, 6H, OC(CF3)2CH3), 1.85 (s, 3H, OC(CF3)2CH3).

13C{1H}

NMR (126 MHz, CDCl3, 25 °C): δ 182.5 (q, OC(CF3)CH, JC-F =

39.1 Hz), 174.0 (q, OC(CF3)CH, JC-F = 37.8 Hz), 122.9 (q, OC(CF3)2CH3, JC-F = 288.5 Hz), 122.8 (q, OC(CF3)2CH3, JC-F = 286.0 Hz), 122.3 (q, OC(CF3)2CH3, JC-F = 288.5 Hz), 117.8 (q, OC(CF3)CH, JC-F = 286.0 Hz), 116.2 (q, OC(CF3)CH, JC-F = 282.2 Hz), 96.6 (s, OC(CF3)CH), 88.6 (sp, OC(CF3)2CH3, JC-F = 31.5 Hz), 86.2 (sp, OC(CF3)2CH3, JC-F = 31.5 Hz), 15.9 (s, OC(CF3)2CH3), 15.6 (s, OC(CF3)2CH3).

19F{13C}

NMR (282 MHz, CDCl3, 25

°C): δ -74.71 (s, 3F, OC(CF3)CH), -76.72 (s, 6F, OC(CF3)2CH3), -77.10 (s, 3F,

ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces 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 59

OC(CF3)CH), -77.29 (m, 6F, OC(CF3)2CH3), -77.47 (m, 6F, OC(CF3)2CH3). Anal. Calc’d for WO6C17H10F24: C: 21.49; H: 1.06%. Found: C: 21.66; H: 0.82%. Thermolysis of 5: NMR analysis. A sample of 5 (0.10 g, 0.11 mmol) was added to a 10 mL Schlenk tube in a nitrogen-filled glove box and capped with a septum. The arm of the Schlenk tube was connected to a U-tube via Tygon tubing. A flow of nitrogen gas was introduced into the Schlenk tube via needle on the Schlenk line. The gas flow rate was measured with a bubbler attached to the other end of the U-tube. A sand bath was pre-heated to 250 °C. Then the Schlenk tube was submerged into the sand bath and the U-tube was placed in a liquid nitrogen bath to collect the thermolysis products for 60 min. After 60 min, the U-tube was removed from the liquid nitrogen bath and allowed to warm to room temperature. CDCl3 (0.50 mL) was used to dissolve the thermolysis products, which were characterized by 1H and 19F NMR. Thermolysis of 5: GC-MS analysis. A sample of 5 (0.10 g, 0.11 mmol) was added to a 10 mL screw-top headspace vial in a nitrogen-filled glove box and capped with a septum lid. The vial was placed into an oven pre-heated to 250 °C for 45 min. A sample

ACS Paragon Plus Environment

8

Page 9 of 59 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

ACS Applied Materials & Interfaces

of the headspace gas was collected using a 10 μL glass syringe and needle and characterized by GC-MS. Volatility Analysis. Sublimation experiments were performed on a Schlenk line equipped with MKS pressure controlling equipment (throttling valve 253A-1-40-2, pressure controller 651CD2S1N, and absolute pressure transductor 121AA-01000B). The active pressure set point during sublimation was 1300 ± 1 mTorr. In a nitrogen-filled glovebox, 100-200 mg of the compound to be studied (1-9 sequentially in individual experiments) was loaded into a 10 mL Schlenk tube equipped with a drip tip cold finger. Prior to heating, the Schlenk tube was placed under a flow of N2 until the cold finger was cooled to 10 °C. The Schlenk tube was placed in a silicone oil heating bath during sublimation. The temperature of the hot plate was ramped from room temperature by 5 °C every 15 minutes until sublimate was detected as the appearance of a white-yellowish coating for 1-6, a yellow coating for 7 and green droplets for 8-9. The reported sublimation temperature was the temperature of the silicone oil bath at the onset of sublimation. AACVD of WOx from 5. Native silicon dioxide (Si/SiOx, n-type, ) and bare glass substrates were prepared by cutting squares of approximately 1 cm2.

The

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces 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 59

substrates were cleaned in boiling acetone, methanol, and deionized water for 3 min each. The substrates were placed on a SiC coated graphite susceptor heated by a radio frequency (RF) induction heat generator within a custom AACVD reactor and held under vacuum for 18 h.

The down-flow vertical impinging jet reactor has been detailed

previously.33,34 Compound 5 (0.912 g, 0.960 mmol) was dissolved in 12 mL of anhydrous diglyme and loaded into a Hamilton 10 mL gastight syringe and pumped into the nebulizer, vibrating at 1.44 MHz, at a rate of 4 mL h-1. High purity nitrogen (99.999 % purity, 1000 sccm) was used as a carrier gas and the reactor pressure was maintained at 650 Torr with a reaction time of 150 min. Conventional CVD of WOx from 5. Conventional depositions were performed on the same reactor as the AACVD depositions. In a nitrogen-filled glovebox, compound 5 (0.50 g) was loaded into a stainless-steel bubbler. High purity nitrogen (99.999 % purity, 1000 sccm) was used as a carrier gas and the reactor pressure was maintained at 650 Torr with a reaction time of 210 min. During deposition, the bubbler was heated in a 90 °C sand bath and the transfer line connecting the outlet of the bubbler to the reaction chamber was heated to 80 °C.

ACS Paragon Plus Environment

10

Page 11 of 59 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

ACS Applied Materials & Interfaces

Materials Characterization. The chemical composition of the deposited materials was studied using X-ray Photoelectron Spectroscopy (XPS) with an ESCALAB 250Xi instrument (Thermo Fisher Scientific, Pittsburgh, PA) working with a monochromatic aluminum anode (hν = 1486.6 eV) as the X-ray source. Survey scans were initially acquired from the as-prepared surface of the deposited film with an electron pass energy of 100 eV and step size of 0.5 eV. High resolution scans for detailed peak analysis were performed at an electron pass energy of 20 eV and an energy step size of 0.1 eV. The binding energies were referenced to the adventitious C 1s peak located at 284.6 eV. New spectra were also acquired after gentle sputtering with 500 eV Ar ions rastered on a 2 mm2 area at various time intervals (30 s to several min). This low energy sputtering was employed to remove surface contaminants while simultaneously minimizing the reduction of tungsten oxide towards metallic W.35

XPS Advantage software was used to

deconvolute the acquired XPS core-level peaks. A Shirley type background line was subtracted before peak fitting. The XPS spectra of the W 4f core level were fitted into peak doublets with parameters of spin-orbit separation ΔEP (4f5/2–4f7/2) = 2.18 eV. The intensity ratio of the W 4f7/2 and W 4f5/2 peak doublet was set to 4:3.48 The O1s peak for

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces 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 59

WO3 films also exhibits components due to oxyhydrates, C-O-C and C-O-H species, as well as adsorbate H2O.48 The crystalline structure of the deposited films and structures was investigated with grazing incidence X-ray diffraction (GIXRD) (Empyrean, Panalytical) working with a Cu anode (45 kV and 40 mA power settings). The instrument was set in a parallel beam geometry with an X-ray mirror and a (1/8)° slit in the incident optics side and a 0.27° parallel beam collimator in front of a detector on the diffraction side. Scans were recorded from 20° to 80° 2θ angles with a 0.03º step size and a 3 s time per step. The incidence angle of X-ray on the samples surface was adjusted within the 1° to 4° range, depending on the film thickness, in order to maximize the diffraction signal from the deposits while reducing the contribution from the Si substrate. The acquired diffraction patterns were fitted using the Panalytical HighscorePlus software to identify the crystalline phases using the ICCD 2018 data base. The grain size was estimated from a line profile analysis using several of the most intense diffraction lines. The morphologies of the deposits were measured by field emission scanning electron microscopy (FESEM, FEI Nova NanoSEM 430).

Results and Discussion

Precursor Design. Single-source precursors bearing terminal oxo, alkoxide, and β-diketonate/β-ketoesterate ligands have been previously reported for the deposition of

ACS Paragon Plus Environment

12

Page 13 of 59 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

ACS Applied Materials & Interfaces

metal oxides.24,36,37 Terminal oxo ligands, either present in the precursor or generated from the alkoxide ligands during decomposition, provide the source of oxygen in the film deposits, eliminating the need for secondary oxygen sources.37 Chelating ligands can improve the stability of the complexes to air and moisture by occupying an empty coordination site, if one is present on the metal. Also, chelating ligands have been proven to be effective for obtaining monomeric complexes, should the alkoxide ligands lack sufficient steric bulk to prevent oligomerization.38 The physical properties necessary for AACVD precursors are different from those for conventional CVD. For conventional CVD, in which neat precursor is volatilized in a bubbler, a moderately high vapor pressure is critical.39 In contrast, the AACVD process involves nebulization of a solution of the precursor, which is then delivered to the substrate as aerosol droplets.

Therefore, AACVD precursors must have sufficient

solubility in the solvent used to generate the aerosol but their volatility requirements are not as stringent. Because AACVD widens the scope of available precursors due to the decreased need for precursor volatility, AACVD can be used to test ligand types to develop sets of precursors without having to deal with the volatility constraints of CVD.

ACS Paragon Plus Environment

13

ACS Applied Materials & Interfaces 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 59

Then modification of the ligands to increase the precursor volatility would allow use of the complexes in conventional bubbler systems as well. Ideally, newly developed precursors would be suitable for use in both deposition methods. A common approach for increasing precursor volatility is use of sterically demanding ligands, which diminish intermolecular interactions.37

Incorporation of

fluoroalkyl substituents on the ligands has also been shown to increase volatility by the introduction of repulsive interactions between non-bonding p electrons on the fluorines.40 We previously reported the series of complexes WO(OR)3L [R = C(CH3)2CF3 and C(CF3)2CH3, L = acetylacetonate (acac), dipivaloylmethanate (dpm), ethyl acetoacetate (etac), and tert-butyl pivaloyl acetate (tbac)], where fluorination was present only on the alkoxide.30 Our interest in testing the effects of fluorination of the chelating ligands suggested complexes 1-9 as target precursors. Synthesis of Precursors.

The oxo-tetraalkoxide complexes WO(OR)4 [R =

C(CH3)2CF3, C(CF3)2CH3, CH(CF3)2] react with one equivalent of the β-diketone or βketoester [HL = Hacac, Hhfac, Hetfac] upon dropwise addition at 0 °C followed by warming to room temperature and stirring for 15 h to form the WO(OR)3L complexes 1-9

ACS Paragon Plus Environment

14

Page 15 of 59 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

ACS Applied Materials & Interfaces

as shown in Scheme 1. The solvent and volatile byproducts are removed in vacuo and 1-9 are purified via sublimation or distillation.

Scheme 1. Synthesis of compounds 1-9.

O

O

R1O W OR1 R1O OR1

O

Et2O

+

0 °C, 15 h

R3

R2

R 1O 1

R O

O W O

OR1 O R3

R2

WO(OR)3L

R1

R2

R3

1

C(CH3)2CF3

CH3

CH3

2

C(CH3)2CF3

CF3

CF3

3

C(CH3)2CF3

OCH2CH3

CF3

4

C(CF3)2CH3

CH3

CH3

5

C(CF3)2CH3

CF3

CF3

6

C(CF3)2CH3

OCH2CH3

CF3

7

CH(CF3)2

CH3

CH3

8

CH(CF3)2

CF3

CF3

9

CH(CF3)2

OCH2CH3

CF3

NMR Characterization. The splitting patterns and intensities of the signals in the 1H, 13C,

19F

and 19F NMR spectra of 1-9 are consistent with their structural assignments. The

NMR spectra for complexes bearing trifluoro-tert-butoxide ligands, 1-3, show two

symmetry inequivalent signals in a 2:1 ratio, corresponding to the CF3 groups. The same pattern is observed in the 1H NMR spectra for the methyl groups on the hexafluoro-tert-

ACS Paragon Plus Environment

15

ACS Applied Materials & Interfaces 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 59

butoxide ligands of 4-6 and for the methine protons on the hexafluoro-isopropoxide ligands of 7-9. In the 13C NMR spectra, the tertiary carbons on the alkoxide ligands for 16, and the secondary carbons for 7-9, also display two signals in a 2:1 ratio which are split by coupling to fluorines two bonds away (2JC-F ≈ 30 Hz).

In the 1H NMR spectra for 1-3, there are three distinct signals attributed to the methyl groups on the alkoxide ligands. Two of the signals arise from two alkoxide ligands that are trans to one another. The methyl groups on these alkoxides are inequivalent with respect to each other but each methyl group has an enantiotopic group located on the other alkoxide. This results in two sets of enantiotopic methyl groups, consistent with the

mer isomer. The last signal is from the third alkoxide bearing equivalent CH3 groups. The analogous enantiotopic groups are also observed for the CF3 groups for compounds 4-6. The integrations in the 1H NMR and 19F NMR spectra for 1-9 confirm that only one β-diketonate or β-ketoesterate is coordinated to the complex. Furthermore, the

ACS Paragon Plus Environment

16

Page 17 of 59 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

ACS Applied Materials & Interfaces

substituents on the β-diketonate ligand display symmetry inequivalence with respect to each other which is further consistent with the mer isomer. Mass Spectrometry. Investigating decomposition mechanisms of CVD precursors enables improvement in precursor design by understanding how these complexes fragment during the CVD process.

One strategy is analyzing mass spectrometry

fragmentation patterns as they are known to model the gas-phase portions of precursor decomposition during CVD.41 However, over-interpretation should be avoided as mass spectrometry produces charged fragments whereas CVD produces neutral fragments.42 Mass spectra were obtained for compounds 2, 3 and 5 using electrospray ionization (ESI) (Figure S32-34, Supporting Information). This softer ionization technique results in less fragmentation. However, the lack of fragmentation can be compensated by coupling ESIMS with tandem mass spectrometry (MS/MS), which allows for further fragmentation of a selected peak using additional collision energy. The observed molecular ion was a sodium adduct of the complex, [M + Na]+. This fragment was further fragmented by MS/MS. Table 1 lists selected ions of interest.

ACS Paragon Plus Environment

17

ACS Applied Materials & Interfaces 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 59

Table 1. Selected observed ions for ESI MS/MS for compounds 2, 3, and 5 in positive mode from [M+Na]+ WO(OR)3L

WO(OC(CH3)2CF3)3(hfac)

[M+Na]+

[M-

[M-C4H5F3-

[M-C4H5F3-

[M-HL+H]+

m/z (%)b

HOR+H]+

HL+H]+

HOR+H]+

m/z (%)

m/z (%)

m/z (%)

m/z (%)

n. o.

471.02 (30)

n. o.

811.04 (6)

581.06

(2) WO(OC(CH3)2CF3)3(etfac)

(100) 787.08 (39)

n. o.

471.02 (83)

527.01 (23)

581.06

(3) WO(OC(CF3)2CH3)3(hfac) (5) an.o.

(100) n.o.a

768.95

n. o.

n. o.

n. o.

(100)

= not observed, bRelative abundances were calculated by normalizing peak

intensities relative to the highest peak at 100 %. The ion that results from loss of C4H3F3, an alkene fragment derived from an alkoxide, and the protonated β-diketonate/β-ketoesterate ligand HL (471 amu) is observed for compounds 2 and 3. This ion corresponds to pathways that involve W-O bond cleavage on the chelate through proton transfer from the alkyl group on a neighboring alkoxide ligand (Scheme 2a). This fragment was also previously observed for 1.30 This common fragment observed for complexes that bear trifluorinated-tertbutoxide ligands (1, 2, and 3) suggests they share similar decomposition pathways. Complex 3 also produced a fragment (527 amu) formed from the loss of C4H5F3 and the fluoroalcohol HOR, both derived from alkoxide ligands. The formation of this ion occurs

ACS Paragon Plus Environment

18

Page 19 of 59 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

ACS Applied Materials & Interfaces

through a proton transfer of a methyl hydrogen on one alkoxide ligand to a neighboring alkoxide (Scheme 2b), which is consistent with the behavior of related alkoxide complexes.43,44 For complex 5, the base peak (768 amu) observed was loss of alcohol by protonation. Fragmentation patterns from the mass spectra of compounds 6-9 were obtained but the fragments could not be identified. Scheme 2. a) Proposed decomposition mechanism to produce the observed ions [MC4H5F3-L+H]+ from 2 and 3, b) Proposed decomposition mechanism to produce the observed ion [M-C4H5F3-HOR+H]+ from 3.

ACS Paragon Plus Environment

19

ACS Applied Materials & Interfaces 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

Thermolysis of 5.

Page 20 of 59

Thermolysis of 5 was performed to obtain mechanistic

information from conditions that more closely resemble CVD. Mass spectrometry detects ionized species whereas both CVD and thermolysis are more likely to produce neutral fragments. A neat sample of 5 was heated above 250 °C and the volatile products were characterized by NMR and GC-MS (Scheme 3). In the 1H and

19F

NMR spectra of the

thermolysis products, the only identifiable signals were from hexafluoroacetylacetone (HL), and HOC(CF3)2CH3 (HOR) (Figures S25-26, Supporting Information). Analyzing the headspace gas by GC-MS after thermolysis resulted in detection of the fluorinated alkene, (CF3)2C=CH2 (164 amu), and HL (208 amu). Although the molecular ion of HOR (182 amu) was not observed directly, ions produced from its fragmentation were observed (Figure S27-28, Supporting Information). All observed spectral fragmentation patterns were compared with NIST library spectra to confirm their identities. The presence of these thermolysis products provides further support that hexafluorinated-tert-butoxide complexes share similar decomposition pathways with their trifluorinated analogues.

ACS Paragon Plus Environment

20

Page 21 of 59 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

ACS Applied Materials & Interfaces

Scheme 3. Products of the thermolysis of 5.

CF3 F 3C

O

F 3C

O CF3

CF3 O W O

O

CF3

O CF3

O



+

> 250 °C

F 3C

CF3

F 3C

CF3

O + HO CF3

CF3

F 3C 5

Thermolysis of 8. Thermolysis of 8 was also performed in the same manner as for 5. A neat sample of 8 was heated above 250 °C for 45 min (Scheme 4). Analysis of the headspace gas by GC-MS resulted in the detection of hexafluoroacetone (166 amu), hexafluoroacetylacetone (208 amu), and hexafluoroisopropanol (99 amu) (Figures S2931, Supporting Information).

The formation of these products is consistent with

intramolecular proton transfer as proposed above for 5.

The formation of

hexafluoroacetone and hexafluoroacetylacetone is attributed to a deprotonation of a methine proton on an alkoxide ligand by the chelating ligand (Scheme 5). The proposed route to hexafluoroacetone and hexafluoroisopropanol involves deprotonation of a methine proton on an alkoxide by a neighboring alkoxide, shown in Scheme 6. Formation of the ketone contrasts with compounds bearing tert-butoxide ligands, which have no

ACS Paragon Plus Environment

21

ACS Applied Materials & Interfaces 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 59

methine hydrogen. In the partially fluorinated tert-butoxide cases, deprotonation of a methyl group results in the cleavage of the W-O bond.

Scheme 4. Products of the thermolysis of 8.

CF3

CF3

F 3C

O

F 3C

O CF3

O

O

W

CF3

O

 F 3C

> 250 °C

CF3

CF3

CF3

O + HO CF3

+

O

O

O F 3C

CF3

F 3C 8

Scheme 5.

Proposed decomposition mechanism to produce the products

hexafluoroacetylacetone and hexafluoroacetone from 8.

O CF3

CF3

F 3C

O

F 3C

O CF3

O W O

O O

H

F 3C

O

O

+ CF3

F 3C

CF3

CF3 WO(OR)2 CF3

F 3C 8

ACS Paragon Plus Environment

22

Page 23 of 59 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

ACS Applied Materials & Interfaces

Scheme 6.

Proposed decomposition mechanism to produce the products

hexafluoroisopropanol and hexafluoroacetylacetone from 8.

CF3 HO CF3 F 3C F 3C

H

CF3

O O

CF3

O W O

O

O +

CF3

CF3

F 3C

CF3

WO(OR)L

O CF3

F 3C 8

Thermogravimetric Analysis.

Thermogravimetric analysis (TGA) was used to

obtain information about the volatilization temperatures and decomposition of 1-9 (Figures 1a-b).

Compounds 2 and 8 undergo decomposition concomitant with

sublimation leaving residual masses of 16.3% and 17.6% respectively. Decomposition occurring simultaneously with sublimation for compounds 2 and 8 suggests compounds that contain the hfac ligand are inherently less stable. However, when there is greater steric bulk and more fluorine on the alkoxide ligands, the compound exhibits greater volatility, allowing clean sublimation as is the case for compound 5. All other compounds

ACS Paragon Plus Environment

23

ACS Applied Materials & Interfaces 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 59

undergo complete sublimation or evaporation without any decomposition at atmospheric pressure, indicating adequate volatility and thermal stability for transport during the CVD process. Complexes with fluorinated β-diketonates (2, 5, and 8) undergo initial mass loss at much lower temperatures in contrast to their non-fluorinated analogues 1, 4 and 7. Similarly, 3 and 6 exhibit evaporation temperatures 20-30 °C lower than their counterparts WO(OC(CH3)2CF3)3(etac) and WO(OC(CF3)2CH3)3(etac), respectively, which are not fluorinated on the chelating ligand.30

This augmented volatility is attributed to the

repulsive nature of the fluorine lone pairs, which lessens intermolecular interactions.45 Once the alkoxide ligand has a CF3 group, increased fluorination on the alkoxide has a negligible effect on volatility as 1 and 4 have similar sublimation temperatures despite 4 exhibiting

greater

fluorine

content

on

the

alkoxide.

Complexes

in

the

hexafluoroisopropoxide series, 7-9, overall exhibit lower sublimation temperatures than 1-4.

ACS Paragon Plus Environment

24

Page 25 of 59 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

ACS Applied Materials & Interfaces

Figure 1. TGA plots of compounds a) 1-5 and b) 6-9.

Volatility Analysis.

Determining the sublimation temperature of a precursor can

be used as a surrogate for measuring its vapor pressure.46 The volatility of 1-9 was thus assessed by sublimation on a Schlenk line at 1300 mTorr (Table 2). This pressure was held constant to within ± 1 mTorr using a throttle valve and digital pressure controller. The onset of sublimation (Tsub) was determined by the oil bath temperature at which sublimate could be detected on a cold finger cooled to 10 °C. The TGA curves for 1-9 (Figures 1a-b) demonstrate that these compounds remain stable when heated to Tsub at

ACS Paragon Plus Environment

25

ACS Applied Materials & Interfaces 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 59

atmospheric pressure. Therefore, it was assumed that no decomposition occurred during sublimation. For compounds that bear tert-butoxide ligands (1-5) there is little change to Tsub when an additional CF3 group is added only on the alkoxide. However, when additional CF3 groups are added onto the chelate, the decrease in Tsub is more pronounced. Fluorination on the chelate thus has a greater effect on volatility than additional fluorination on the alkoxide. Compounds bearing hexafluoro-isoproxide ligands (7-9) exhibited lower sublimation temperatures than the analogous compounds with tertbutoxide ligands. Lastly, the sufficient volatility and thermal stability demonstrated for these compounds permits volatilization in a bubbler, thus enabling a comparison between materials grown by CVD and AACVD. Table 2. Temperatures for the onset of sublimation for 1-9.

WO(OR)3L

Tsub (°C) at 1300 ±1 mTorr

WO(OC(CH3)2CF3)3(acac) (1)

48

WO(OC(CH3)2CF3)3(etfac) (3)

47a

WO(OC(CH3)2CF3)3(hfac) (2)

35

ACS Paragon Plus Environment

26

Page 27 of 59 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

ACS Applied Materials & Interfaces

WO(OC(CF3)2CH3)3(acac) (4)

47

WO(OC(CF3)2CH3)3(etfac) (6)

43a

WO(OC(CF3)2CH3)3(hfac) (5)

32

WO(OCH(CF3)2)3(acac) (7)

40

WO(OCH(CF3)2)3(etfac) (9)

38a

WO(OCH(CF3)2)3(hfac) (8)

34a

aEvaporation

of liquid precursor, not sublimation

Growth of WOx from AACVD and Conventional CVD. Compound 5 exhibited the lowest onset sublimation temperature of 32 °C and was therefore chosen as the precursor for deposition studies. For AACVD, precursor 5 was volatilized from a 0.080 M solution in diglyme. For CVD, 0.50 g of neat precursor was loaded into a stainless-steel bubbler for direct volatilization.

Growth of tungsten oxide on silicon substrates was

achieved at deposition temperatures of 350, 400, 450 and 500 °C. A blue color was observed for the deposited films and became more apparent with increasing deposition temperature.

Scanning Electron Microscopy. Scanning electron microscopy (SEM) was used to investigate the morphology of the deposits. As the deposition temperature decreased,

ACS Paragon Plus Environment

27

ACS Applied Materials & Interfaces 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 28 of 59

the coverage of the film also decreased and areas of the substrate could be observed. In AACVD (Figure 2a), deposits grown at 400 and 450 °C displayed a platelet-like morphology with more dense growth observed at 450 °C. Grain sizes for deposits grown at 350, 400, and 450 °C appear to be 140 – 160 nm (Figure S44). Nanorods or nanowires were observed in samples grown at 500 °C with diameters measuring at 250 – 350 nm. Bundled nanorods have been previously observed for deposits grown from similar tungsten oxo-alkoxide complexes at higher temperatures.47

In CVD (Figure 2b),

nanocubes were deposited at temperatures above 350 °C. Nanocube sizes remained about the same for deposits grown at 350 and 400 °C (60 – 80 nm) and increased at 500 °C (100 – 250 nm) (Figure S45). Complete coverage of the substrate was achieved only at 500 °C. An amorphous deposit, as determined by GIXRD, was obtained at 350 °C.

ACS Paragon Plus Environment

28

Page 29 of 59 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

ACS Applied Materials & Interfaces

Figure 2. Plane-view SEM images of deposits grown on Si at 500, 450, 400 and 350 °C from precursor 5 by a) AACVD and b) CVD.

Grazing Incidence X-Ray Diffraction. Grazing incidence X-ray diffraction (GIXRD) was used to determine the crystallinity of the deposited material (Figure 3a-b). The deposits grown by CVD at 350 °C were found to be almost amorphous. At higher growth temperatures, strong diffraction peaks were observed, indicative of highly crystalline material. In the CVD deposits, the most intense peaks at around 23° 2θ were assigned to the (002), (020), and (200) planes of monoclinic WO3 (ICDD 04-007-1277). Other monoclinic and orthorhombic WO3 phases were also matched to the recorded patterns

ACS Paragon Plus Environment

29

ACS Applied Materials & Interfaces 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 59

with similar matching scores. The diffraction patterns for the AACVD deposits grown at 450, 400 and 350 °C also match well with the same crystal structure as the CVD deposits. However, at 500 °C, the diffraction peaks were consistent with monoclinic WO2.83 phase (ICDD 00-036-0103) or orthorhombic WO2.63 phase (ICCD 04-005-5745). Cubic WO3 phases (ICCD 00-046-1096 and 00-041-0905) could also match the strong diffraction lines near 23° 2θ, but the primary crystal structure should either be a monoclinic or orthorhombic phase in order to account for all of the observed peaks. The broadening of the diffraction peaks could be due to the presence of multiple phases of monoclinic, orthorhombic and cubic WO3, with overlapping peaks.

Grain sizes were estimated from line profile analysis to be 12 – 83 nm. These grain sizes are smaller in comparison to the measurements obtained from SEM images. The larger grain sizes from SEM images are most likely due to clusters of several grains which have slightly different crystalline orientations.

ACS Paragon Plus Environment

30

Page 31 of 59 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

ACS Applied Materials & Interfaces

Figure 3. Grazing incidence X-ray diffraction patterns for deposits grown from precursor 5 by a) CVD and b) AACVD with labeled diffraction peaks corresponding to monoclinic WO3 (violet), monoclinic WO2.83 (green), and orthorhombic WO2.63 (blue).

X-Ray Photoelectron Spectroscopy. X-ray photoelectron spectroscopy (XPS) was used to analyze the elemental composition and the chemical bonding of the deposited films. A survey of the as-deposited films revealed that fluorine was not present, indicating that the fluorinated groups dissociate cleanly from the precursor (Figure S36, Supporting Information). The core-level spectra for as-deposited W 4f and O 1s for both CVD and AACVD are shown in Figures 4a-b and 5a-b. The deconvoluted W 4f spectrum showed

ACS Paragon Plus Environment

31

ACS Applied Materials & Interfaces 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 59

the presence of primarily W6+ with minor sub-stoichiometric W5+ (Figure S41, Supporting Information). The W 4f7/2 binding energies (BE) for W6+ (35.5 eV) and W5+ (34.0 eV) were similar among the films deposited by both methods across all temperatures. Gentle sputtering was used to remove any surface-bound intermediates. Adventitious carbon, with a BE of 284.6 eV, was observed on the material surface and its amount was reduced by sputtering (Figures S37-38, Supporting Information).

As expected, continued

sputtering of the sample preferentially removes oxygen atoms from the material, reducing the W and producing additional W 4f7/2 peaks with BE of 33.3 eV for W4+ and 31.6 eV for metallic W (Figure S42, Supporting Information). The preferential sputtering of O in tungsten oxide has been previously reported in the literature.35 As a result, accurate O:W ratios cannot be obtained from either as-deposited or sputtered samples. However, as determined by GIXRD, the crystalline material deposited was WO3 or very close to this stoichiometry. The presence of WO3 was also confirmed by Rutherford backscattering investigations for deposits that presented a smooth surface morphology (Figure S43, Supporting Information).

ACS Paragon Plus Environment

32

Page 33 of 59 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

ACS Applied Materials & Interfaces

The O 1s peak structure was rather complex. In addition to O2- in the WO3 lattice (530.6 eV), there was also the presence of hydroxides, WOxHy (531.2 eV), C-O-C, C-OH compounds (~532.5 eV) and adsorbed H2O (533.1 eV).48 The relative fractions of these components depended on the surface structure and morphology. After gentle sputtering, the peaks corresponding to water, hydroxides and carbonites decreased in intensity while the peak corresponding to WO3 increased. Continued sputtering resulted in further broadening of the O 1s peak, indicative of the formation of several new types of tungsten oxides due to the reduction of WO3 from Ar ion sputtering (Figure S39-S40, Supporting Information).35 For films deposited at lower substrate temperatures, Si from the substrate appeared in the survey spectra, which corroborated the SEM images that indicated that the deposits do not completely cover the substrate (Figure 2).

ACS Paragon Plus Environment

33

ACS Applied Materials & Interfaces 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 34 of 59

Figure 4. W 4f7/2 and W f5/2 core-level spectra for a) AACVD and b) CVD deposits grown from 5 at 500, 450, 400, and 350 °C.

Figure 5. O 1s core-level spectra for a) AACVD and b) CVD deposits grown from 5 at 500, 450, 400, and 350 °C.

ACS Paragon Plus Environment

34

Page 35 of 59 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

ACS Applied Materials & Interfaces

Conclusions

Tungsten oxo complexes bearing partially fluorinated tert-butoxide and βdiketonate/ketoesterate ligands were synthesized as precursors for CVD and AACVD of tungsten oxide.

Assessment of their thermal properties was performed by

thermogravimetric analysis and determination of sublimation temperatures. Complexes 1, 3-5, 6-7, and 9 sublimed cleanly without any decomposition while 2 and 8 exhibited decomposition concomitant with sublimation.

Mass spectrometry of 3 showed

decomposition pathways that lead to the loss of fluoroalcohol (HOC4H7F3) and trifluoroisobutene (C4H5F3) or loss of β-diketonate (Hhfac) and trifluoroisobutene, with both pathways occurring through proton transfer and cleavage of a C-O bond. Compound 5 was selected as the precursor for AACVD and CVD deposition studies and tungsten oxide was deposited at 500, 450, 400 and 350 °C. Although fully stoichiometric WO3 was deposited by both deposition methods (as determined by GIXRD and XPS), there were some notable differences among the deposits. GIXRD showed that all crystalline samples were monoclinic WO3, except for the sample grown at 500 °C

ACS Paragon Plus Environment

35

ACS Applied Materials & Interfaces 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 36 of 59

by AACVD which was assigned to a mixture of orthorhombic-monoclinic phases. Scanning electron microscopy images showed differences in morphology between materials grown by the two deposition methods. In AACVD deposits, nanowires were grown at 500 °C and nanoplates were grown at 450 and 400 °C whereas in CVD deposits, nanocubes were grown at those same temperatures. This difference in morphology suggests the solvent used in AACVD plays a role during film growth. Having the ability to deposit material by both deposition methods from the same precursor allows for not only flexibility in choice of delivery method, but also in the ability to obtain a broadened range of film properties. Further studies probing differences between CVD and AACVD are underway.

ASSOCIATED CONTENT Supporting Information. Experimental procedures, NMR spectra for 2-9, mass spectra for 2-9, thermolysis analysis for 5 and 8, particle size distribution histograms for CVD and AACVD deposits, RBS data for WOx deposition of 5, XPS of WOx deposition of 5.

AUTHOR INFORMATION

ACS Paragon Plus Environment

36

Page 37 of 59 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

ACS Applied Materials & Interfaces

Corresponding Author

*Email: [email protected]

ACKNOWLEDGMENTS

We thank the National Science Foundation for partial support of this work under grant number CHE-1213965 and thank the Laplas Nucleu project for funding.

Mass

spectrometry was performed at the Mass Spectrometry Research and Education Center at University of Florida, supported by grant NIH S10 OD021758-01A1. We also thank Dr. Dan Pantelica and Mrs. Diana Maria Dracea (IFIN-HH, Romania) for the RBS analysis.

ACS Paragon Plus Environment

37

ACS Applied Materials & Interfaces 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 38 of 59

REFERENCES

(1) Cong, S.; Geng, F.; Zhao, Z. Tungsten Oxide Materials for Optoelectronic Applications, Adv Mater 2016, 28, 10518-10528. (2) More, A. J.; Patil, R. S.; Dalavi, D. S.; Suryawanshi, M. P.; Burungale, V. V.; Kim, J. H.; Patil, P. S. Synthesis and Characterization of Potentiostatically Electrodeposited Tungsten Oxide Thin Films for Smart Window Application, J.

Electron. Mater. 2016, 46, 974-981. (3) Szilágyi, I. M.; Saukko, S.; Mizsei, J.; Tóth, A. L.; Madarász, J.; Pokol, G. Gas Sensing Selectivity of Hexagonal and Monoclinic WO3 to H2S, Solid State Sci. 2010, 12, 1857-1860. (4) Tsai, S. Y.; Ni, C. T.; Fung, K. Z. Characterization of Infrared Transmittance in Mixed Transition Metal Oxides for Solar Cells Application, Ceram. Int. 2017, 43, S460-S463.

ACS Paragon Plus Environment

38

Page 39 of 59 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

ACS Applied Materials & Interfaces

(5) Gao, L.; Gan, W.; Qiu, Z.; Zhan, X.; Qiang, T.; Li, J. Preparation of Heterostructured WO3/TiO2 Catalysts from Wood Fibers and its Versatile Photodegradation Abilities, Sci. Rep. 2017, 7, 1102. (6) Late, D. J.; Kashid, R. V.; Rout, C. S.; More, M. A.; Joag, D. S. Low Threshold Field Electron Emission from Solvothermally Synthesized WO2.72 Nanowires,

Appl. Phys. A: Mater. Sci. Process. 2010, 98, 751-756. (7) Ramana, C. V.; Utsunomiya, S.; Ewing, R. C.; Julien, C. M.; Becker, U. Structural Stability and Phase transitions in WO3 Thin Films, J. Phys. Chem. B 2006, 110, 10430-10435. (8) Žumer, M.; Nemanič, V.; Zajec, B.; Wang, M.; Wang, J.; Liu, Y.; Peng, L.-M. The Field-Emission and Current−Voltage Characteristics of Individual W5O14 Nanowires, J. Phys. Chem. C 2008, 112, 5250-5253. (9) Hu, W. B.; Zhu, Y. Q.; Hsu, W. K.; Chang, B. H.; Terrones, M.; Grobert, N.; Terrones, H.; Hare, J. P.; Kroto, H. W.; Walton, D. R. M. Generation of Hollow Crystalline Tungsten Oxide Fibres, Appl. Phys. A: Mater. Sci. Process. 2000, 70, 231-233.

ACS Paragon Plus Environment

39

ACS Applied Materials & Interfaces 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 40 of 59

(10) Wang, H.; Quan, X.; Zhang, Y.; Chen, S. Direct Growth and Photoelectrochemical Properties of Tungsten Oxide Nanobelt Arrays,

Nanotechnology 2008, 19, 065704. (11) Choi, H. G.; Jung, Y. H.; Kim, D. K. Solvothermal Synthesis of Tungsten Oxide Nanorod/Nanowire/Nanosheet, J. Am. Ceram. Soc. 2005, 88, 1684-1686. (12) Hsieh, Y.-T.; Chang, L.-W.; Chang, C.-C.; Shih, H. C. Synthesis of WO3 Nanorods by Thermal CVD at Various Gas Flow Rates and Substrate Temperatures, Electrochem. Solid-State Lett. 2011, 14. (13) Barreca, D.; Bozza, S.; Carta, G.; Rossetto, G.; Tondello, E.; Zanella, P. Structural and Morphological Analyses of Tungsten Oxide Nanophasic Thin Films Obtained by MOCVD, Surf. Sci. 2003, 532-535, 439-443. (14) Choy, K. L. Chemical Vapour Deposition of Coatings, Prog. Mater. Sci. 2003, 48, 57-170. (15) Bekermann, D.; Barreca, D.; Gasparotto, A.; Maccato, C. Multi-component Oxide Nanosystems by Chemical Vapor Deposition and Related Routes: Challenges and Perspectives, CrystEngComm 2012, 14.

ACS Paragon Plus Environment

40

Page 41 of 59 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

ACS Applied Materials & Interfaces

(16) Hitchcock, P. B.; Pye, P. L. Crystal and Molecular Structure of Chloro(triphenyl phosphite)gold(I), J. Chem. Soc., Dalton Trans. 1977, 1457-1460. (17) Riaz, U. Low-temperature Atmospheric-Pressure Chemical Vapor Deposition of Tungsten Oxide Thin Films, Thin Solid Films 1993, 235, 15-16. (18) Blackman, C. S.; Correig, X.; Katko, V.; Mozalev, A.; Parkin, I. P.; Alcubilla, R.; Trifonov, T. Templated Growth of Tungsten Oxide Micro/Nanostructures Using Aerosol Assisted Chemical Vapor Deposition, Mater. Lett. 2008, 62, 4582-4584. (19) Powell, M. J.; Carmalt, C. J. Aerosols: A Sustainable Route to Functional Materials, Chemistry 2017, 23, 15543-15552. (20) Marchand, P.; Hassan, I. A.; Parkin, I. P.; Carmalt, C. J. Aerosol-Assisted Delivery of Precursors for Chemical Vapour Deposition: Expanding the Scope of CVD for Materials Fabrication, Dalton Trans. 2013, 42, 9406-9422. (21) Nagai, M.; Kishida, K.; Omi, S. Low Pressure Thermal CVD Synthesis of Tungsten Nitride Thin Film using WCl6 as Tungsten Source, Nippon Kagaku

Kaishi 1996, 368-374.

ACS Paragon Plus Environment

41

ACS Applied Materials & Interfaces 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 42 of 59

(22) Yoon, H. A.; Yang, M. X.; Xi, M., Deposition of Tungsten Films on Oxide Films from W(CO)6, 2001, Patent US6218301B1 (23) Wang, X.; Lin, J.; Zhu, Y.; Luo, C.; Suenaga, K.; Cai, C.; Xie, L. Chemical Vapor Deposition of Trigonal Prismatic NbS2 Monolayers and 3R-Polytype Few-Layers,

Nanoscale 2017, 9, 16607-16611. (24) Mäkelä, M.; Hatanpää, T.; Mizohata, K.; Räisänen, J.; Ritala, M.; Leskelä, M. Thermal Atomic Layer Deposition of Continuous and Highly Conducting Gold Thin Films, Chem. Mater. 2017, 29, 6130-6136. (25) Basato, M.; Brescacin, E.; Tondello, E. Amorphous WO3 Films via "Wet" CVD of a W(VI) Oxoalkoxide Precursor, Chem. Vap. Deposition 2001, 7, 219-224. (26) Cross, W. B.; Parkin, I. P.; O'Neill, S. A.; Williams, P. A.; Mahon, M. F.; Molloy, K. C. Tungsten Oxide Coatings from the Aerosol-Assisted Chemical Vapor Deposition of W(OAr)6 (Ar = C6H5, C6H4F-4, C6H3F2-3,4); Photocatalytically Active -WO3 Films, Chem. Mater. 2003, 15, 2786-2796. (27) Kim, H.; Bonsu, R. O.; Bock, D. C.; Ou, N. C.; Korotkov, R. Y.; McElwee-White, L.; Anderson, T. J. Tungsten Oxide Film and Nanorods Grown by Aerosol-

ACS Paragon Plus Environment

42

Page 43 of 59 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

ACS Applied Materials & Interfaces

Assisted Chemical Vapor Deposition using κ2-β-Diketonate and β-Ketoesterate Tungsten (VI) Oxo-Alkoxide Precursors, ECS Journal of Solid State Science and

Technology, 2016, 5, Q3095-Q3105. (28) Bonsu, R. O.; Bock, D. C.; Kim, H.; Korotkov, R. Y.; Abboud, K. A.; Anderson, T. J.; McElwee-White, L. Synthesis and Evaluation of 2--Diketonate and Ketoesterate Tungsten(VI) Oxo-Alkoxide Complexes as Precursors for Chemical Vapor Deposition of WOx Thin Films, Dalton Trans. 2016, 45, 10897-10908. (29) Doppelt, P. Why is Coordination Chemistry Stretching the Limits of MicroElectronics Technology?, Coord. Chem. Rev. 1998, 178-180, 1785-1809. (30) Bock, D. C.; Ou, N. C.; Bonsu, R. O.; Anghel, C. T.; Su, X.; McElwee-White, L. Synthesis of Tungsten Oxo Fluoroalkoxide Complexes WO(OR)3L as Precursors for Growth of WOx Nanomaterials by Aerosol-Assisted Chemical Vapor Deposition, Solid State Ionics 2018, 315, 77-84. (31) Vernardou, D.; Drosos, H.; Spanakis, E.; Koudoumas, E.; Katsarakis, N.; Pemble, M. E. Electrochemical Properties of Amorphous WO3 Coatings Grown

ACS Paragon Plus Environment

43

ACS Applied Materials & Interfaces 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 44 of 59

on Polycarbonate by Aerosol-Assisted CVD, Electrochim. Acta 2012, 65, 185189. (32) Bhachu, D. S.; Waugh, M. R.; Zeissler, K.; Branford, W. R.; Parkin, I. P. Textured Fluorine-Doped Tin Dioxide Films Formed by Chemical Vapour Deposition,

Chemistry 2011, 17, 11613-11621. (33) Bchir, O. J.; Johnston, S. W.; Cuadra, A. C.; Anderson, T. J.; Ortiz, C. G.; Brooks, B. C.; Powell, D. H.; McElwee-White, L. MOCVD of Tungsten Nitride (WNx) Thin Films from the Imido Complex Cl4(CH3CN)W(NiPr), J. Cryst. Growth 2003, 249, 262-274. (34) McClain, K. R.; O'Donohue, C.; Shi, Z.; Walker, A. V.; Abboud, K. A.; Anderson, T.; McElwee-White, L. Synthesis of WN(NMe2)3 as a Precursor for the Deposition of WNx Nanospheres, Eur. J. Inorg. Chem. 2012, 4579-4584. (35) Xie, F. Y.; Gong, L.; Liu, X.; Tao, Y. T.; Zhang, W. H.; Chen, S. H.; Meng, H.; Chen, J. XPS Studies on Surface Reduction of Tungsten Oxide Nanowire Film by Ar+ Bombardment, J. Electron Spectrosc. Relat. Phenom. 2012, 185, 112-118.

ACS Paragon Plus Environment

44

Page 45 of 59 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

ACS Applied Materials & Interfaces

(36) Bloor, L. G.; Carmalt, C. J.; Pugh, D. Single-Source Precursors to Gallium and Indium Oxide Thin Films, Coord. Chem. Rev. 2011, 255, 1293-1318. (37) Jones, A. C.; Aspinall, H. C.; Chalker, P. R. Molecular Design of Improved Precursors for the MOCVD of Oxides Used in Microelectronics, Surf. Coat. Tech. 2007, 201, 9046-9054. (38) Clegg, W.; Errington, J. R.; Kraxner, P.; Redshaw, C. Solid State and Solution Studies of Tungsten (VI) Oxotetraalkoxides, J. Chem. Soc. Dalton Trans. 1992, 1431 - 1438. (39) Jones, A. C.; Hitchman, M. L. Chemical Vapor Deposition : Precursors, Processes and Applications; RSC Publishing: 2008, p 1-36. (40) Miinea, A. L.; Suh, S.; Hoffman, M. D. Indium Fluoroalkoxide Compounds, Inorg.

Chem. 1999, 38, 4447-4454. (41) Won, Y. S.; Kim, Y. S.; Anderson, T. J.; McElwee-White, L. Computational Study of the Gas Phase Reactions of Isopropylimido and Allylimido Tungsten Precursors for Chemical Vapor Deposition of Tungsten Carbonitride Films: Implications for the Choice of Carrier Gas, Chem. Mater. 2008, 20, 7246-7251.

ACS Paragon Plus Environment

45

ACS Applied Materials & Interfaces 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 46 of 59

(42) Won, Y. S.; Kim, Y. S.; Anderson, T. J.; Reitfort, L. L.; Ghiviriga, I.; McElweeWhite, L. Homogeneous Decomposition of Aryl- and Alkylimido Precursors for the Chemical Vapor Deposition of Tungsten Nitride:  A Combined Density Functional Theory and Experimental Study, J. Am. Chem. Soc. 2006, 128, 13781-13788. (43) Bonsu, R. O.; Kim, H.; O'Donohue, C.; Korotkov, R. Y.; Abboud, K. A.; Anderson, T. J.; McElwee-White, L. Dioxo-Fluoroalkoxide Tungsten(VI) Complexes for Growth of WOx Thin Films by Aerosol-Assisted Chemical Vapor Deposition,

Inorg. Chem. 2015, 54, 7536-7547. (44) Terry, K. W.; Ganzel, P. K.; Tilley, T. D. Low-Temperature Pyrolytic Transformations of Tri-tert-Butoxysiloxy Derivatives of Aluminum to Aluminosilicate Materials, Chem. Mater. 1992, 4, 1290-1295. (45) Buchanan, W. D.; Guino-O, M. A.; Ruhlandt-Senge, K. Highly Volatile Alkaline Earth Metal Fluoroalkoxides, Inorg. Chem. 2010, 49, 7144-7155. (46) Carden, W.; Pedziwiatr, J.; Abboud, K. A.; McElwee-White, L. Halide Effects on the Sublimation Temperature of X-Au-L Complexes: Implications for Their Use as

ACS Paragon Plus Environment

46

Page 47 of 59 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

ACS Applied Materials & Interfaces

Precursors in Vapor Phase Deposition Methods, ACS Appl. Mater. Interfaces 2017, 9, 40998-41005. (47) Bonsu, R. O.; Kim, H.; O'Donohue, C.; Korotkov, R. Y.; McClain, K. R.; Abboud, K. A.; Ellsworth, A. A.; Walker, A. V.; Anderson, T. J.; McElwee-White, L. Partially Fluorinated Oxo-Alkoxide Tungsten(VI) Complexes as Precursors for Deposition of WOx Nanomaterials, Dalton Trans. 2014, 43, 9226-9233. (48) Shpak, A. P.; Korduban, A. M.; Medvedskij, M. M.; Kandyba, V. O. XPS Studies of Active Elements Surface of Gas Sensors Based on WO3−x Nanoparticles, J.

Electron Spectrosc. Relat. Phenom. 2007, 156-158, 172-175.

ACS Paragon Plus Environment

47

ACS Applied Materials & Interfaces 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

Scheme 1. Synthesis of compounds 1-9. 161x155mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 48 of 59

Page 49 of 59 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

ACS Applied Materials & Interfaces

Scheme 2. a) Proposed decomposition mechanism to produce the observed ions [M-C4H5F3-L+H]+ from 2 and 3, b) Proposed decomposition mechanism to produce the observed ion [M-C4H5F3-HOR+H]+ from 3. 236x151mm (150 x 150 DPI)

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 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

Scheme 3. Products of the thermolysis of 5. 200x52mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 50 of 59

Page 51 of 59 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

ACS Applied Materials & Interfaces

Scheme 4. Products of the thermolysis of 8. 208x52mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 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

Scheme 5. Proposed decomposition mechanism to produce the products hexafluoroacetylacetone and hexafluoroacetone from 8. 163x61mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 52 of 59

Page 53 of 59 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

ACS Applied Materials & Interfaces

Scheme 6. Proposed decomposition mechanism to produce the products hexafluoroisopropanol and hexafluoroacetylacetone from 8. 161x64mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 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

Figure 1. TGA plots of compounds a) 1-5 and b) 6-9. 356x151mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 54 of 59

Page 55 of 59 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

ACS Applied Materials & Interfaces

Figure 2. Plane-view SEM images of deposits grown on Si at 500, 450, 400 and 350 °C from precursor 5 by a) AACVD and b) CVD. 356x151mm (150 x 150 DPI)

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 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

Grazing incidence X-ray diffraction patterns for deposits grown from precursor 5 by a) CVD and b) AACVD with labeled diffraction peaks corresponding to monoclinic WO3 (violet), monoclinic WO2.83 (green), and orthorhombic WO2.63 (blue).

ACS Paragon Plus Environment

Page 56 of 59

Page 57 of 59 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

ACS Applied Materials & Interfaces

Figure 4. W 4f7/2 and W f5/2 core-level spectra for a) AACVD and b) CVD deposits grown from 5 at 500, 450, 400, and 350 °C. 710x315mm (72 x 72 DPI)

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 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

Figure 5. O 1s core-level spectra for a) AACVD and b) CVD deposits grown from 5 at 500, 450, 400, and 350 °C.

ACS Paragon Plus Environment

Page 58 of 59

Page 59 of 59 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

ACS Applied Materials & Interfaces

TOC graphics 264x264mm (96 x 96 DPI)

ACS Paragon Plus Environment