Reaction Probability and Infrared Detection of the ... - ACS Publications

Erin Durke Davis‡; Alec Wagner†; Monica McEntee§; Manpreet Kaur†; Diego Troya†; John R. Morris*†. View Author Information. † Department o...
0 downloads 0 Views 3MB Size
Subscriber access provided by UniSA Library

Letter

Reaction Probability and Infrared Detection of the Primary Ozonide in Collisions of O with Surface-bound C 3

60

Erin Durke Davis, Alec Wagner, Monica McEntee, Manpreet Kaur, Diego Troya, and John R. Morris J. Phys. Chem. Lett., Just Accepted Manuscript • Publication Date (Web): 08 Oct 2012 Downloaded from http://pubs.acs.org on October 11, 2012

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 free 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 accessible to all readers and 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.

The Journal of Physical Chemistry Letters 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 20

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

The Journal of Physical Chemistry Letters

Reaction Probability and Infrared Detection of the Primary Ozonide in Collisions of O3 with Surface-bound C60 Erin Durke Davis ([email protected]),* Alec Wagner ([email protected]), Monica McEntee ([email protected]),** Manpreet Kaur ([email protected]), Diego Troya ([email protected]), and John R. Morris ([email protected]) Department of Chemistry, Virginia Tech, Blacksburg, Virginia 24061 *US Army, Edgewood Chemical Biological Center, Edgewood, MD, 21010. ** Department of Chemistry, University of Virginia, Charlottesville, VA, 22908

The kinetics and mechanism of reactions between gas-phase ozone and surface-bound C60 have been investigated by monitoring the changes to in situ reflection-absorption infrared spectra within a well-characterized film of C60 during exposure to a controlled flux of pure ozone. These ultrahigh vacuum studies provide direct infrared spectroscopic evidence for the formation and decomposition of a primary ozonide of C60. The spectral assignments of this highly unstable intermediate have been verified using electronic structure calculations. Theory and experiment revealed that C60 is oxidized nearly exclusively via addition of ozone across the double bond that links two 6-carbon containing rings of the molecule. Following spectral characterization, the initial probability for ozone to react with the surface was found to be 5.8 ± 0.2 × 10-4. Once formed, the ozonide quickly thermally decomposed to a variety of carbonylcontaining products.

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

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

TOC Image

Keywords: ozonide, ozone, fullerene, oxidation, infrared, mechanism

2

ACS Paragon Plus Environment

Page 2 of 20

Page 3 of 20

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

The Journal of Physical Chemistry Letters

Since their discovery in 1985,1 fullerenes have been explored for and implemented in a diverse array of practical applications including new organometallics,2,

3

electronics,4-6

medicine,7-9 and in the field of surface science.10, 11 Today, the most common fullerene, C60, is produced by the metric ton,12 which has sparked concerns over the possible impact the molecule may have on the environment. Once C60 enters the environment, this large polycyclic aromatic molecule likely reacts efficiently with highly oxidizing atmospheric gases such as O3, OH, and NOx compounds. However, there is currently little known about the reaction kinetics and mechanisms of C60 with these gas-phase molecules. Therefore, we have explored the initial reaction channels involved in the oxidation of surface-bound C60 by gas-phase O3. In contrast to experiments at the gas-surface interface, condensed-phase reactions between C60 and ozone have been explored by a number of groups. For example, recent research indicates that ozone readily oxidizes C60 in various solvents, including water.13-30 In 2000, Heymann et al. detected and isolated an unstable intermediate as a result of the ozonation of C60/toluene solutions. These researchers employed liquid chromatographic methods to help identify the intermediate as C60O3, the primary ozonide (PO) of C60.26 However, definitive spectroscopic characterization of the PO has yet to be reported. Similar to C60, the PO of alkenes and other polycyclic aromatic hydrocarbons is known to be extremely short-lived. The short lifetime of the PO under most experimental conditions has precluded its direct spectroscopic detection at ambient temperatures. Therefore, several studies of the reaction between ozone and alkenes have been conducted in matrix isolation type experiments at low temperatures.31-33 These studies have proven helpful in providing infrared spectroscopic characterization of the PO for a variety of unsaturated organic molecules.31-33 Far fewer studies of the reaction between gas-phase ozone and solid or vaporous C60 have been conducted.34, 35 Most relevant to the study described below, Cataldo et al.36 explored the oxidation of solid powder C60 with gas-phase ozone. They reported the observation of a stable reaction product, which they identified to be the primary ozonide of C60. In addition, these studies proposed that the rate of PO formation followed first-order kinetics characterized by a rate constant of k = 2.0 x 10-4 s-1.36

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

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

In this Letter, we present the results of an ultrahigh vacuum (UHV) surface science study into the reaction between a well-controlled flux of pure O3 and surface-bound C60. By tracking the formation of new bonds on the surface with vibrational spectroscopy during ozone exposure, we have determined the rate of emergence of the PO of C60. In addition, with the aid of electronic structure calculations, we have characterized the infrared spectrum of this short-lived intermediate at the gas-surface interface, which has provided detailed insight into the dynamics that govern the initial stages of the reaction. The surface samples were created in vacuum by vapor deposition of C60 from a temperature-controlled Knudsen cell onto a clean polycrystalline Au surface. The Au substrate served as the support for the C60 molecules as well as a reflective mirror to facilitate reflectionabsorption infrared spectroscopy (RAIRS). Prior to these experiments, a quartz crystal microbalance (QCM) was employed to directly measure the coverage of C60 on the surface during film formation. For the study reported below, the film thickness was 50 nm, which, according to previous work,37 corresponds to approximately 40 ML of fullerenes above the underlying gold substrate. The QCM measurements showed that, under our experimental conditions and range of surface coverages, the RAIRS absorbance for the fullerene films was directly proportional to the surface concentration. Therefore, the changes in intensity of the main C60 modes during ozone exposure were used to reveal the extent of oxidation and track the kinetics of the reaction. Once the Au surface was coated with a thin film of fullerene molecules, the sample was transferred directly into the main UHV chamber (base pressure of < 10-9 Torr) without exposure to atmosphere. In the UHV system, pure ozone was dosed directly onto the sample via a capillary array doser coupled to an ozone storage and purification system.38, 39 The concentration of ozone was periodically measured using an in-line UV-Vis spectrometer. In addition, the relative flux of ozone into the chamber was continually monitored with a quadrupole mass spectrometer tuned to the parent ion for O3 (m/z = 48). The final flux of ozone on the surface was estimated by measuring the UV absorbance of the O3 gas source at 254 nm immediately prior to entering the capillary array doser. This absorbance was converted to a concentration using Beers Law, and the flux through the doser was determined from a calculation based on the physical dimensions

4

ACS Paragon Plus Environment

Page 4 of 20

Page 5 of 20

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

The Journal of Physical Chemistry Letters

of the doser and the kinetic theory of gases. The calculated value for flux correlated well with the pressure rise in the vacuum chamber during O3 exposure to a clean, Au surface. The reaction between C60 and ozone was monitored in situ with RAIRS. RAIRS was enabled by a Bruker IFS 66v/S spectrometer, which was coupled to the UHV chamber via a set of KBr windows on either side of the sample in a highly glancing configuration. The IR radiation was detected via a liquid nitrogen cooled mercury cadmium telluride (MCT) detector. Each individual spectrum shown in this report represents the average of 100 scans with 2 cm-1 resolution. A clean Au surface was used as a background, unless otherwise indicated. Figure 1 A shows the infrared spectrum of the as-prepared C60 surface sample. Two of the four IR-active modes for C60 appear in the spectrum, at 1182 and 1428 cm-1.40 The other two IR-active modes, at 577 and 527 cm-1,40 are outside the detection range of the MCT detector (4000-750 cm-1). The smaller peaks present in this spectrum are due to combination,

13

C

isotopes, higher-order IR, and Raman modes of the C60 layer adjacent to the gold surface that become IR active due to disruption of the molecular symmetry of the adsorbate.40 Following initial characterization, the C60-covered Au surface was exposed to a continuous source of ozone and changes to the film were tracked with RAIRS. Figure 1 B displays a subset of the difference spectra recorded continuously during exposure. For the difference spectra, the spectrum of the original C60 sample (Figure 1 A) served as the background such that negative peaks represent modes that decreased in intensity and positive features correspond to surface species that emerged during reaction. The difference spectra clearly show that the intensities of the major fullerene peaks at 1182 and 1428 cm-1 were reduced immediately upon exposure to ozone. As these peaks decreased in intensity, several new modes appeared. The spectrum recorded following 400 L (1 L = 1 Langmuir = 10-6 Torr · s) of exposure, labeled c in Figure 1 B, shows the emergence of six new modes that developed in the low-wavenumber region of the spectrum. The intensity of these new modes reached a maximum at this stage in the experiment before the peaks returned to baseline after 4 × 104 L of exposure. As these lowwavenumber modes decreased, they were replaced by the final products, which appear to be a multi-component surface-bound carbonyl species.

5

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

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. C60 film on polycrystalline Au surface. IR active C60 modes shown at 1182 and 1428 cm-1. (A) Initial C60 spectrum, before ozone exposure. (B) Difference spectra recorded during the ozone exposure of the C60 surface in UHV. The amount of ozone exposed to the surface is indicated in Langmuir. Spectrum b shows the simultaneous decrease of the main C60 modes at 1182 and 1428 cm-1, and the appearance of the six new modes in the lowwavenumber region. After approximately 400 L of ozone exposure, the six new modes have reached maximum intensity and begin to decrease, until they disappear entirely following 4 × 104 L of exposure.

It should be noted that the extent of reaction appeared to be restricted to the top-most layers of the C60 film. That is, we found the rate of reaction and the final IR intensity for the product peaks to be insensitive to overall film thickness and the majority of the C60 molecules were unaffected by the ozone. For the experiment highlighted in Fig. 1, the main IR peaks for the fullerene molecules decreased by only 20% following extensive ozone exposure. Furthermore, the IR modes for the fullerene molecules immediately adjacent to the Au surface were unaffected by ozone exposure, which indicates that the underlying Au surface likely does not contribute to the observed surface chemistry. These observations are consistent with the work of Cataldo et al., which showed that only the outer-most layer of a solid C60 sample was oxidized during extensive ozonolysis.36 The broad carbonyl feature that emerged in the IR spectra (see spectra d-f in Figure 1 B) may be due to the highly oxidized and cross-linked species known as ozopolymer, which likely contains several different types of carboxy-functionalized products.19,20,36,41 Characterization of these complex infrared features has yet to be accomplished, but is the subject of on-going research in our group. For the present study, we focus on the intermediate species characterized by a transient, but intense, peak at 977 cm-1 and accompanied by at least five additional infrared bands.

6

ACS Paragon Plus Environment

Page 6 of 20

Page 7 of 20

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

The Journal of Physical Chemistry Letters

We hypothesize that the infrared transient at 977 cm-1, along with the five additional modes, is due to the formation of the primary ozonide of C60. The Criegee mechanism for ozonolysis of olefins suggests that O3 first adds across a double bond to form the PO.42, 43 The PO is well documented to be highly unstable and thermally decomposes to form a diradical and, eventually, carbonyl species. In fact, direct detection of the PO for any ozone-surface reaction has proven to be difficult due to the short-lived nature of this intermediate, and consequently little precedence exists in the literature for definitive assignment of the vibrational spectra in Figure 1 B. In order to aid in the assignment of the IR spectral features measured in this work, we have performed electronic structure calculations of the two symmetry-inequivalent primary ozonides resulting from addition of ozone to a bond shared between two hexagons (6-6) or between a hexagon and a pentagon (6-5) in C60. The calculated harmonic frequencies were computed after full geometry optimization of the PO of C60 (Cs symmetry) at the B3LYP/6-31G* level, and were scaled by the recommended factor of 0.96.44 A subset of additional calculations at the M06/6-31G* level provided results consistent with the ones reported here at the B3LYP/631G* level. All calculations were conducted with the Gaussian09 suite of programs. Figure 2 displays spectrum c from Figure 1 B, the time during the measurements at which the intermediate species reached maximum intensity, along with the calculated harmonic vibrational frequencies for the primary ozonide of C60 corresponding to the addition of O3 to a 66 bond in C60 (Figure 2 A) and to the addition of O3 to a 6-5 bond in C60 (Figure 2 B). The remarkable agreement between the theoretically calculated vibrational spectrum for the PO of C60 on a 6-6 bond and the experimental IR spectrum (typical differences are within 20 cm-1) verifies that these modes are, in fact, due to the formation of the PO intermediate.

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

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 2. The experimental spectrum displayed in both graphs, A and B, is the difference spectrum (c) from Figure 1 B. In graph A, the spectrum is overlaid with the calculated vibrational frequencies of the PO resulting from ozone addition to a 6-6 bond. Graph B compares the spectrum with the calculated vibrational frequencies of a PO formed through ozone addition across a 6-5 bond.

Comparison of the experimental spectrum to the calculations performed on the PO resulting from the addition of O3 to a 6-5 bond in C60 (Figure 2 B) does not produce the level of agreement found for the addition to a 6-6 bond (Figure 2 A). This result, together with the fact that the calculated barrier for O3 addition to a 6-5 bond is 4.3 kcal/mol, but the addition to a 6-6 bond is barrierless (the transition-state energy is actually below the reagents’ asymptote for the 6-6 addition), clearly indicates that the experimental spectra are dominated by the PO formed at a 6-6 bond. Our calculated barriers are in agreement with the minimum-energy-pathway calculations of Sabirov et al.45 Because the C60 film is continuously exposed to ozone in these studies and the addition of ozone to the 6-6 bonds is an energetically barrierless process, it is likely that multiple molecules add to a single C60. Calculations carried out for all possible isomers that result from the addition of two ozone molecules to different 6-6 bonds provide IR spectra that are remarkably similar to those of the single-PO spectrum shown in Fig. 2 A; therefore, the calculations provide no direct insight into the number of POs that form on a single C60 molecule. Additional studies, perhaps involving more detailed molecular dynamics simulations, will be required to fully elucidate the extent of oxidation on any particular fullerene surface molecule.

8

ACS Paragon Plus Environment

Page 8 of 20

Page 9 of 20

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

The Journal of Physical Chemistry Letters

In addition to characterizing the IR spectrum of the PO, we have studied rate of formation of this intermediate to determine the initial reaction probability for ozone impinging on the fullerene-covered surface. Figure 3 displays a plot of the change in absorbance for the main C60 mode as a function of time, along with the increase in the intensity of the most intense mode associated with the PO of C60 (977 cm-1). As highlighted in this figure, the observed rate constants for the time evolution of these two modes are identical within our experimental uncertainty, which indicates that the formation of the intermediate is the first step toward the overall decomposition of the C60 molecules and

Figure 3. The absorbance of the primary fullerene peak at 1428 cm-1 and of the intermediate peak at 977 cm-1 as a function of time during O3 exposure.

subsequent creation of a complex array of carbonyl-containing functional groups. The initial reaction probability, γ0, which is simply the ratio of reactive collisions to total collisions, can be extracted from the rate constant for the formation of the PO according to:46

γ0 =

# of "reactive"collisions kobs = # of total collisions Aprod ΦO3

(1),

where Aprod is the surface area occupied by the intermediate product, the PO, and ΦO3 is the flux of ozone. A non-linear least squares analysis of the data in Fig. 3 provides the observed rate constant (kobs in Eqn. 1), which was used to determine an initial reaction probability of γ0 = 5.8 ± 0.2 x 10-4. The value for γ0 reported here is surprisingly similar to the rate at which ozone reacts in interfacial encounters with other unsaturated organics,47-54 despite the fact that the current study is among the few to explore this reaction in the absence of background gases and on highly characterized clean surfaces. Notably, Thornberry and Abbatt55 utilized a coated flow tube, coupled with a chemical ionization mass spectrometer, to measure the reaction probabilities for

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

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 20

ozone with oleic acid, linoleic acid, and linolenic acid. The values for γo in their work were 8.0 x 10-4, 1.3 x 10-3, and 1.8 x 10-3 for the three acids, respectively.55 These values are within a factor of three of the experimentally determined probability for ozone reacting with surface-bound fullerene molecules. Perhaps slightly more relevant to the current study, are the recent UHVbased experiments reported by Fiegland and co-workers,49 and Lu et al.,47 which show that the reactivity of ozone with a well-organized alkene-terminated self-assembled monolayer is much lower than that for reactions with the disordered acid films employed in the work of Abbatt et al.54 These differences may be due to the steric constraints present in an organized monolayer that limit the ability of the ozone molecule to reach a favorable configuration for addition to the double bond. Similar constraints may play a somewhat less important role in the chemistry of ozone on fullerene films reported here. Overall, there are many competing factors that play significant roles in determining the rate and initial reaction probability of the ozone-fullerene reaction, including the energetics of PO formation on the surface of the highly polycyclic aromatic molecules, the steric and structural aspects of PO formation on a surface of closely-packed fullerene molecules, and the possible structural changes to the surface that continually alter the dynamics of the initial O3surface collision. Subsequent theoretical and experimental studies will be used to help elucidate the most important contributions to the reaction rate in this complex system. These studies will lead to a better understanding of the oxidation chemistry of C60 and other carbonaceous nanomaterials that may ultimately affect their overall hydrophobicity and environmental transport.

10

ACS Paragon Plus Environment

Page 11 of 20

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

The Journal of Physical Chemistry Letters

Experimental Details The fullerene molecules used in this study were purchased from Acros Organics (99.99%) and used without further purification. The temperature-controlled Knudsen cell was designed and constructed by McAllister Technical Services. Polycrystalline Au surfaces employed in the work were obtained from EMF Corporation. They were composed of 1000 Å of Au deposited on a Cr-coated silica surface. The UV-Vis spectrometer (model: EPP2000) employed for characterization of the ozone source gas was acquired from StellarNet, Inc.. The gas phase reactants and products for these reactions were monitored by a quadrupole mass spectrometer (Extrel CMS, LLC).

Acknowledgements The authors would like to thank the NSF for supporting this work (CHE-0948293). We are also grateful to Professor Linsey Marr and Andrea Tiwari for many helpful discussions. References 1. Kroto, H. W.; Heath, J. R.; Obrien, S. C.; Curl, R. F.; Smalley, R. E., C-60 Buckminsterfullerene. Nature 1985, 318 (6042), 162-163. 2. Sawamura, M.; Kuninobu, Y.; Toganoh, M.; Matsuo, Y.; Yamanaka, M.; Nakamura, E., Hybrid of ferrocene and fullerene. Journal of the American Chemical Society 2002, 124 (32), 9354-9355. 3. Toganoh, M.; Matsuo, Y.; Nakamura, E., Synthesis of ferrocene/hydrofullerene hybrid and functionalized bucky ferrocenes. Journal of the American Chemical Society 2003, 125 (46), 13974-13975. 4. Paci, P.; Cappelluti, E.; Grimaldi, C.; Pietronero, L.; Strassler, S., Nonadiabatic high-T-c superconductivity in hole-doped fullerenes. Physical Review B 2004, 69 (2), -. 5. Varshney, D., Pairing mechanism and superconductivity of Rb3C60 fullerides. Journal of Superconductivity 2000, 13 (1), 171-179. 6. Strobel, P.; Ristein, J.; Ley, L., Ozone-Mediated Polymerization of Fullerene and Fluorofullerene Thin Films. J. Phys. Chem. C 2010, 114, 4317-4323. 7. Huang, S. T.; Liao, J. S.; Fang, H. W.; Lin, C. M., Synthesis and anti-inflammation evaluation of new C-60 fulleropyrrolidines bearing biologically active xanthine. Bioorganic & Medicinal Chemistry Letters 2008, 18 (1), 99-103. 8. Ryan, J. J.; Bateman, H. R.; Stover, A.; Gomez, G.; Norton, S. K.; Zhao, W.; Schwartz, L. B.; Lenk, R.; Kepley, C. L., Fullerene nanomaterials inhibit the allergic response. Journal of Immunology 2007, 179 (1), 665-672. 9. Yamakoshi, Y.; Umezawa, N.; Ryu, A.; Arakane, K.; Miyata, N.; Goda, Y.; Masumizu, T.; Nagano, T., Active oxygen species generated from photoexcited fullerene (C-60) as potential medicines: O-2(-center dot) versus O-1(2). Journal of the American Chemical Society 2003, 125 (42), 12803-12809. 10. Swami, N. S. Surface Science Studies of Fullerene and Fulleride Thin Films: Electronic, Vibrational, and Electrical Characterization. Univeristy of Southern California, 1998. 11. Satterley, C. J. P.; A., L. M.; Saywell, A.; Magnano, G.; Rienzo, A.; Mayor, L. C.; Dhanak, V. R.; Beton, P. H.; O'Shea, J. N., Electrospray deposition of fullerenes in ultrahigh vacuum: in situ scanning

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

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

tunneling microscopy and photoemission spectroscopy. Nanotechnology 2007, 18 (45), 455304/1455304/5. 12. Colvin, V. L., The potential environmental impact of engineered nanomaterials. Nature Biotechnology 2003, 21 (10), 1166-1170. 13. Anachkov, M. P.; Cataldo, F.; Deneva, M. G.; Rakovski, S. K., Study of the ozonolysis rate constants' dependence from the degree of oxidation of C60 fullerene and trans-1,4-polychloroprene. Fullerenes, Nanotubes, and Carbon Nanostructures 2007, 15 (3), 181-193. 14. Anachkov, M. P.; Cataldo, F.; Rakovsky, S. K., Reaction kinetics of C60 fullerene ozonation. Fullerenes, Nanotubes, and Carbon Nanostructures 2003, 11 (2), 95-103. 15. Anachkov, M. P.; Cataldo, F.; Rakovsky, S. K., Ozone Reaction with C70 and C60 Fullerenes: The Effect of Temperature on the Reaction Kinetics. Fullerenes, Nanotubes, and Carbon Nanostructures 2004, 12 (4), 745-752. 16. Bulgakov, R. G.; Akhmadieva, R. G.; Musavirova, A. S.; Abdrakhmanov, A. M.; Ushakova, Z. I.; Sharifullina, F. M., The first example of chemiluminescence of fullerenes - oxidation of C60 by ozone in solution. Russian Chemical Bulletin (Translation of Izvestiya Akademii Nauk, Seriya Khimicheskaya) 1999, 48 (6), 1190. 17. Bulgakov, R. G.; Musavirova, A. S.; Abdrakhmanov, A. M.; Nevyadovskii, E. Y.; Khursan, S. L.; Razumovskii, S. D., Chemiluminescence in Ozonolysis of Solutions of Fullerene C60. Journal of Applied Spectroscopy (Translation of Zhurnal Prikladnoi Spektroskopii) 2002, 69 (2), 220-225. 18. Bulgakov, R. G.; Nevyadovskii, E. Y.; Belyaeva, A. S.; Golikova, M. T.; Ushakova, Z. I.; Ponomareva, Y. G.; Dzhemilev, U. M.; Razumovskii, S. D.; Valyamova, F. G., Water-soluble polyketones and esters as the main stable products of ozonolysis of fullerene C60 solutions. Russian Chemical Bulletin (Translation of Izvestiya Akademii Nauk, Seriya Khimicheskaya) 2004, 53 (1), 148159. 19. Cataldo, F., Polymeric fullerene oxide (fullerene ozopolymers) produced by prolonged ozonation of C60 and C70 fullerenes. Carbon 2002, 40 (9), 1457-1467. 20. Cataldo, F.; Ori, O., Ozone reaction with C60 fullerene. A study on the antiozonant activity of C60 fullerene in dienic rubber. Polymer Degradation and Stability 1995, 48 (2), 291-6. 21. Davydov, V. Y.; Filatova, G. N.; Knipovich, O. M.; Lunin, V. V., Study of the kinetics of fullerenes ozonolysis. Proceedings - Electrochemical Society 1996, 96-10 (Recent Advances in the Chemistry and Physics of Fullerenes and Related Materials, Vol. 3), 1295-1306. 22. Deng, J.-P.; Mou, C.-Y.; Han, C.-C., Mass spectrometric study of C60O and isomers of C60O2. Proceedings - Electrochemical Society 1995, 95-10 (Proceedings of the Symposium on Recent Advances in the Chemistry and Physics of Fullerenes and Related Materials, 1995), 1409-24. 23. Deng, J.-P.; Mou, C.-Y.; Han, C.-C., Electrospray and Laser Desorption Ionization Studies of C60O and Isomers of C60O2. Journal of Physical Chemistry 1995, 99 (41), 14907-10. 24. Fortner, J. D.; Kim, D. I.; Boyd, A. M.; Falkner, J. C.; Moran, S.; Colvin, V. L.; Hughes, J. B.; Kim, J. H., Reaction of Water-Stable C-60 aggregates with ozone. Environmental Science & Technology 2007, 41 (21), 7497-7502. 25. Hao, C.; Wu, S.; Teng, Q.-W.; Zhao, X.-Z.; Li, W.-L.; Tang, A.-Q.; Feng, J.-K., Spectroscopic study on the isomers of C60O. Huaxue Xuebao 1997, 55 (6), 622-624. 26. Heymann, D.; Bachilo, S. M.; Weisman, R. B.; Cataldo, F.; Fokkens, R. H.; Nibbering, N. M. M.; Vis, R. D.; Chibante, L. P. F., C60O3, a Fullerene Ozonide: Synthesis and Dissociation to C60O and O2. Journal of the American Chemical Society 2000, 122 (46), 11473-11479. 27. Heymann, D.; Chibante, L. P. F., Reaction of fullerenes C60 and C70 with ozone at different temperatures. Recueil des Travaux Chimiques des Pays-Bas 1993, 112 (10), 531-4. 28. Li, X.; Cai, R.-f.; Kong, Q.-y.; Zhao, L.; Li, Y.-f.; Yu, H.-k.; Hou, J.; Hou, H.-q., Studies on oxidation of C60. Gaodeng Xuexiao Huaxue Xuebao 2001, 22 (4), 515-519. 29. Mi, W.-h.; Liu, Y.-j.; Hao, C., Structure and spectra of C60O3 isomers. Dalian Ligong Daxue Xuebao 2005, 45 (2), 172-176.

12

ACS Paragon Plus Environment

Page 12 of 20

Page 13 of 20

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

The Journal of Physical Chemistry Letters

30. Razumovskii, S. D.; Bulgakov, R. G.; Nevyadovskii, E. Y., Kinetics and Stoichiometry of the Reaction of Ozone with Fullerene C60 in a CCl4 Solution. Kinetics and Catalysis (Translation of Kinetika i Kataliz) 2003, 44 (2), 229-232. 31. Feltham, E. J.; Almond, M. J.; Marston, G.; Ly, V. P.; Wiltshire, K. S., Reactions of alkenes with ozone in the gas phase: a matrix-isolation study of secondary ozonides and carbonyl-containing products. Spectrochimica Acta Part A 2000, 56, 2605-2616. 32. Horie, O.; Moortgat, G. K., Decomposition Pathways of the Excited Criegee Intermediates in the Ozonolysis of Simple Alkenes. Atmospheric Environment Part a-General Topics 1991, 25 (9), 18811896. 33. Coleman, B. E.; Ault, B. S., Investigation of the thermal and photochemical reactions of ozone with 2,3-dimethyl-2-butene Journal of Physical Chemistry A 2010, 114 (48), 12667-74. 34. McElvany, S. W.; Holliman, C. L., Gas-phase reactions of fullerenes ions with ozone. Proceedings - Electrochemical Society 1996, 96-10 (Recent Advances in the Chemistry and Physics of Fullerenes and Related Materials, Vol. 3), 121-133. 35. McElvany, S. W.; Holliman, C. L., Gas-phase oxidation of fullerene and metallofullerene ions. Proceedings - Electrochemical Society 1997, 97-14 (Recent Advances in the Chemistry and Physics of Fullerenes and Related Materials), 772-782. 36. Cataldo, F., Ozone reaction with carbon nanostructures 1: reaction between solid C60 and C70 fullerenes and ozone. Journal of Nanoscience and Nanotechnology 2007, 7 (4/5), 1439-1445. 37. Tang, L.; Zhang, X.; Guo, Q., Organizing C60 molecules on a nanostructured Au(111) surface. Surface Science 2010, 604 (15-16), 1310-1314. 38. Yates, J. T., Jr., Experimental Innovations in Surface Science. 1 ed.; AIP/Springer-Verlag: New York, 1998; p 920. 39. Zhukov, V.; Popova, I.; Yates, J. T., Jr., Delivery of pure ozone in ultrahigh vacuum. Journal of Vacuum Science & Technology A 2000, 18 (3). 40. Wang, K. A.; Rao, A. M.; Eklund, P. C.; Dresselhaus, M. S.; Dresselhaus, G., Observation of Higher-Order Infrared Modes in Solid C60 Films. Physical Review B 1993, 48 (15), 11375-11380. 41. Cataldo, F.; Heymann, D., A study of polymeric products formed by C60 and C70 fullerene ozonation. Polymer Degradation and Stability 2000, 70 (2), 237-243. 42. Criegee, R.; Schroder, G., A crystallized primary ozonide. Chemische Berichte 1960, 93 (3), 689700. 43. Criegee, R.; Wenner, G., The ozonation of 9,10-oktalins. Justus Liebigs Annalen der Chemie 1949, 564 (1), 9-15. 44. http://webbook.nist.gov/chemistry/ 45. Sabirov, D. S.; Khursan, S. L.; Bulgakov, R. G., Ozone addition to C-60 and C-70 fullerenes: A DFT study. Journal of Molecular Graphics & Modelling 2008, 27 (2), 124-130. 46. Paz, Y.; Trakhtenberg, S.; Naaman, R. J., Reaction between O(3P) and Organized Organic Thin Films. J. Phys. Chem. 1994, 98 (51), 13517-13523. 47. Lu, J. W.; Fiegland, L. R.; Davis, E. D.; Alexander, W. A.; Wagner, A.; Gandour, R. D.; Morris, J. R., Initial Reaction Probability and Dynamics of Ozone Collisions with a Vinyl-Terminated SelfAssembled Monolayer Journal of physical Chemistry C 2011, 115 (51), 25343-25350. 48. Dubowski, Y.; Vieceli, J.; Tobias, D. J.; Gomez, A.; Lin, A.; Nizkorodov, S. A.; McIntire, T. M.; Finlayson-Pitts, B. J., Interaction of Gas-Phase Ozone at 296K with Unsaturated Self-Assembled Monolayers: A New Look at an Old System. J. Phys. Chem. A 2004, 108, 10473-10485. 49. Fiegland, L. R.; McCorn Saint Fleur, M.; Morris, J. R., Reactions of C=C-Terminated SelfAssembled Monolayers with Gas-Phase Ozone. Langmuir 2005, 21, 2660-2661. 50. Sullivan, R. C.; Thornberry, T.; Abbatt, J. P. D., Atmos. Chem. Phys. 2004, 4, 1301. 51. Chang, R. Y.-W.; Sullivan, R. C.; Abbatt, J. P. D., Initial uptake of ozone on Saharan dust at atmospheric relative humidities. Geophys. Res. Lett. 2005, 32, L14185. 52. Hanisch, F.; Crowley, J. N., Ozone decomposition on Saharan dust: an experimental investigation. Atmos. Chem. Phys. 2003, 3 (1), 119.

13

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters

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

53. Kwamena, N.-O. A.; Thornton, J. A.; Abbatt, J. P. D., Kinetics of Surface-Bound Benzo[a]pyrene and Ozone on Solid Organic and Salt Aerosols. J. Phys. Chem. A 2004, 108 (52), 11626. 54. Kwamena, N.-O. A.; Staikova, M. G.; Donaldson, D. J.; George, I. J.; Abbatt, J. P. D., Role of the Aerosol Substrate in the Heterogeneous Ozonation Reactions of Surface-Bound PAHs. J. Phys. Chem. A 2007, 111 (43), 11050. 55. Thornberry, T.; Abbatt, J. P. D., Heterogeneous reaction of ozone with liquid unsaturated fatty acids: detailed kinetics and gas-phase product studies Physical Chemistry Chemical Physics 2004, 6, 8493.

14

ACS Paragon Plus Environment

Page 14 of 20

Page 15 of 20

0.001

4 B Letters (f) 4 x 10 The L Journal of Physical Chemistry

O3 (Ei, θi)

4

(e)Impulsive 2 x 10 L

O3 Exposure

Δ Absorbance

1 2 3 4 5 6 7 8 9 10 11 12 Direct 13 Reaction 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

Scattering

Desorption

3

(d) 5 x 10 L 2

(c) 4 x 10 L

Accommodation (b) 1 x 10 L 2

(a) 0 L

2000

1500

-1

Wavenumber (cm )

1000

ACS Paragon Plus Environment

L-H Reaction

The Journal of Physical Chemistry Page 16 Letters of A 20

Absorbance

1428cm

-1

1182cm

1 0.005 2 3 4 5 6 7 8 9 ACS Paragon Plus Environment 10 2500 2000 1500 1000 11 -1 Wavenumber (cm ) 12

-1

4 B Page The 17 of (f)Journal 4 x 20 10 L of Physical Chemistry Letters

0.001

4

(e) 2 x 10 L

O3 Exposure

Δ Absorbance

1 2 3 4 5 6 7 8 9 10 11 12

3

(d) 5 x 10 L 2

(c) 4 x 10 L 2

(b) 1 x 10 L (a) 0 L

ACS Paragon Plus Environment 2000

1500

-1

Wavenumber (cm )

1000

Page 18 of 20

A

Δ Absorbance

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22

The Journal of Physical Chemistry Letters

1600

80 60 40 20

Plus Environment 1400ACS Paragon 1200 1000

Wavenumber (cm-1)

800

0

Intensity/ arb. units

0.001

0.001

B

40

Δ Absorbance

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22

The Journal of Physical Chemistry Letters

1600

20

1400 1200 1000 ACS Paragon Plus Environment

Wavenumber (cm-1)

800

Intensity/ arb. units

Page 19 of 20

0

!

The Journal of Physical Chemistry Page 20 Letters 0.002 -2 of -1 20

Integrated Δ Abs.

kobs = 1.03 (± 0.13) x 10 s

1 2 3 4 5 6 7 8 9 10 11

1428 cm-1 977 cm-1

-2

-1

kobs = 1.21Plus (± 0.05) x 10 s ACS Paragon Environment

0

100

200

300

400

Exposure Time (s)

500